Plasmonic Nanomaterials: Characterization and Applications in Organic Synthesis and Catalysis [1 ed.] 9815129090, 9789815129090

This book comprehensively explores the field of plasmonic nanomaterials and their significant impact on organic synthesi

133 109

English Pages 156 Year 2024

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Plasmonic Nanomaterials: Characterization and Applications in Organic Synthesis and Catalysis [1 ed.]
 9815129090, 9789815129090

Citation preview

Plasmonic Nanomaterials

Plasmonic Nanomaterials Characterization and Applications in Organic Synthesis and Catalysis

edited by

Clémence Queffélec D. Andrew Knight

Published by Jenny Stanford Publishing Pte. Ltd. 101 Thomson Road #06-01, United Square Singapore 307591

Email: [email protected] Web: www.jennystanford.com British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library.

Plasmonic Nanomaterials: Characterization and Applications in Organic Synthesis and Catalysis Copyright © 2024 by Jenny Stanford Publishing Pte. Ltd. All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means, electronic or mechanical, including photocopying, recording or any information storage and retrieval system now known or to be invented, without written permission from the publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not required from the publisher. Cover image courtesy: VR2Planets

ISBN 978-981-5129-09-0 (Hardcover) ISBN 978-1-003-47406-7 (eBook)

Contents

Preface 1. Introduction to Plasmonics Wafa Safar, Mathieu Edely, and Marc Lamy de la Chapelle 1.1 Dielectric Constant and Optical Properties of Materials 1.2 Volume Plasmons 1.3 Surface Plasmons 1.3.1 Plasmon Dispersion Curve 1.3.2 Surface Plasmon Characteristics 1.3.3 Surface Plasmon Excitation 1.3.3.1 Plasmon excitation using a grating 1.3.3.2 Plasmon excitation using a prism 1.4 Localized Surface Plasmon 1.4.1 LSP in Spherical Nanoparticles 1.4.2 LSP in Ellipsoidal Nanoparticle 1.4.3 Influence of the Environment 1.4.4 Influence of Metal Nature 1.5 Molecular Plasmonics

2. Characterization of Plasmonic Nanomaterial Gennaro Picardi, Mathieu Edely, and Marc Lamy de la Chapelle 2.1 Ultraviolet–Visible Spectroscopy 2.2 Scanning Electron Microscopy and Transmission Electron Microscopy 2.2.1 Scanning Electron Microscopy 2.2.2 Transmission Electron Microscopy 2.3 Energy-Dispersive X-Ray Spectroscopy

ix

1

2 5 5 5 8 9 9

10 11 13 18 19 21 22 29

29 31 31 32 34

vi

Contents



2.4 2.5

Electron Energy Loss Spectroscopy Atomic Force Microscopy

3. Surface Plasmon Resonance‒Mediated Organic Synthesis Léa Gimeno, Clémence Queffélec, and D. Andrew Knight 3.1 Introduction 3.2 Oxidation Reactions 3.2.1 Oxidation of Alcohols, Aldehydes, and Amines 3.2.2 Oxidation of Alkanes, Alkenes, and Aromatics 3.2.3 Oxidation of Carbon Monoxide 3.3 Reduction Reactions 3.3.1 Reduction of Nitro-Containing Compounds to Azo-Containing Compounds 3.3.2 Reduction of Nitro-Containing Compounds to Amines 3.3.3 Reduction of Aldehydes and Ketones 3.3.4 Miscellaneous Reduction Reactions 3.4 Photothermal Reactions 3.5 Miscellaneous Chemical Reactions 3.5.1 Suzuki–Miyaura Reaction 3.5.2 Photoconversion of CO2 3.5.3 Other Chemical Reactions

4. Plasmon-Mediated Homogeneous Catalysis Léa Gimeno, Somayeh Talebzadeh, Scott Trammell, Clémence Queffélec, and D. Andrew Knight 4.1 Introduction 4.2 Catalytic Hydrolysis 4.2.1 Hydrolysis of Phosphate Esters 4.3 Catalytic Carbon–Carbon Bond Formation 4.3.1 Ullmann Reaction 4.3.2 Henry’s Reaction

34 36 49

49 50 50

57 58 59 59

62 65 65 68 71 71 76 77 97

97 99 99 101 101 104

Contents

5. SERS and TERS and Their Applications in Organic Synthesis or Catalysis Gennaro Picardi, Mathieu Edely, and Marc Lamy de la Chapelle 5.1 Surface-Enhanced Raman Spectroscopy 5.1.1 Electromagnetic Effect 5.1.2 Chemical Effect 5.2 Tip-Enhanced Raman Spectroscopy 5.3 SERS/TERS Applications in Plasmon-Mediated Chemical Reactions

Index

111

111 113 118 123 132 149

vii

Preface

In recent years, the field of nanoscience has witnessed remarkable progress, leading to unprecedented advancements in various scientific disciplines. Among these, plasmonic nanomaterials have emerged as a cornerstone in the fields of materials science and catalysis. Their unique optical properties and surface-enhanced reactivity have revolutionized our approach to organic synthesis and catalytic processes. This book stands at the forefront of this exciting and rapidly evolving field. It is designed to serve as a comprehensive guide for researchers, students, and professionals seeking a deeper understanding of the principles underlying the characterization and utilization of plasmonic nanomaterials in the context of organic synthesis and catalysis. The book provides a thorough exploration of the fundamental principles governing the behavior of plasmonic nanomaterials, coupled with practical insights into their characterization techniques and applications in catalytic processes. Through a meticulous blend of theoretical concepts and experimental methodologies, this book strives to equip readers with the knowledge and skills necessary to harness the full potential of these extraordinary materials. From the basics of plasmonics to advanced characterization techniques, from the design principles of plasmonic catalysts to their applications in diverse organic transformations, this book endeavors to be a valuable resource for both novices entering the field as well as researchers looking to expand their expertise. We are indebted to the dedicated scientists and researchers whose pioneering work in the field of plasmonic nanomaterials has paved the way for the content presented in this book. Their groundbreaking contributions have not only enriched our understanding but have also opened up new frontiers in the domain of catalysis. We hope this book will initiate a wide-ranging discussion about the potential and challenges associated with the use of plasmonic nanomaterials for catalysis. It is our sincere belief that the knowledge

x

Preface

contained within these pages will inspire and empower the next generation of scientists and engineers to push the boundaries of what is possible in the field of organic synthesis and catalysis.

Clémence Queffélec D. Andrew Knight Autumn 2023

Chapter 1

Introduction to Plasmonics

Wafa Safar, Mathieu Edely, and Marc Lamy de la Chapelle

Le Mans University, Laboratoire IMMM, UMR6283, Avenue Olivier Messiaen 72085 Le Mans Cedex 9, France [email protected]

A plasmon is an optical characteristic of a metal resulting from the excitation of free electrons by their interaction with photons or electrons [1, 2]. As these free electrons are weakly bounded to the metal atom, the plasmon corresponds to a collective oscillation of the charge density, named “plasma oscillation.” It occurs when photons or electrons transfer their energy to the free electrons of the metal since their momentum is the same as that of the plasmon. Depending on the geometrical dimensions of the metallic materials, plasmons can be defined and classified as follows:

∑ Volume plasmon (3D plasmon): The plasmon is excited in the material volume. ∑ Delocalized surface plasmon (DSP), also known as surface plasmon polariton (SPP) (2D plasmon): The plasmon is formed at the interface between a dielectric and a metal.

Plasmonic Nanomaterials: Characterization and Applications in Organic Synthesis and Catalysis Edited by Clémence Queffélec and D. Andrew Knight Copyright © 2024 Jenny Stanford Publishing Pte. Ltd. ISBN 978-981-5129-09-0 (Hardcover), 978-1-003-47406-7 (eBook) www.jennystanford.com

2

Introduction to Plasmonics



Having an evanescent character, it can only propagate along the surface. ∑ Localized surface plasmon (LSP) (0D plasmon): The plasmon is confined inside a nanostructure. It can be seen as an eigen mode of oscillation of the free electrons inside a limited volume. Different plasmon modes can be excited depending on the size, shape, and chemical nature of the nanostructure.

1.1 Dielectric Constant and Optical Properties of Materials

To understand plasmons and their features, it is valuable to describe the relation between the optical and electronic properties of a metal. The interaction between the light and the plasmon corresponds to the optical part, whereas the electronic properties refer to the description of the behavior of the free electrons located in the conduction bands of the metal. Several models have been proposed to simulate and explain the electronic properties and their relations with the material. They are based on the definition of the interactions of the electrons with other metal elements, such as the ion array, as well as on the nature of the electrons. Three different models have been proposed: 1. Drude model (1900): The electrons interact only with the metallic ions through collisions [3, 4]. 2. Drude–Lorentz model (1905): Seen as an extension of the Drude model, the electrons are also subject to a restoring force through its electrostatic interaction with the metallic ions [5, 6]. 3. Drude–Sommerfeld model: As the electron is a fermion and a quantic object, it is submitted to the Fermi–Dirac distribution, which has a consequence on its behavior.

In the following, we will present only the Drude–Lorentz model, which is enough to describe the plasmon behavior in the electromagnetic formalism and the interaction between an electromagnetic wave and a material [5, 6].

Dielectric Constant and Optical Properties of Materials

The material is then described by its optical properties, particularly its dielectric constant e. It depends on the excitation frequency and determines the metal’s response to an electromagnetic radiation:

e = e '+ ie " (1.1)



FD = −

The real part of e represents the effect of the electrical field on the polarization of the material. The imaginary part represents the damping of the radiation and the losses within the material. In 1900, Drude first proposed to describe electrons as a gas (the Drude model). The free electrons do not interact with each other but can interact with the atom’s nucleus considered as ions. This interaction between the ions and the electrons is defined as the probability of collision, which induces the damping of the electron motion and is modeled thanks to the viscosity force FD defined by:

m ν t

(1.2)

where t is the average time between two collisions (defined as relaxation time), ν is the electron velocity, and m is the electron mass. Lorentz later refined this model (the Drude–Lorentz model) in 1905 by introducing the electrostatic interaction of the electrons with the positively charged ions corresponding to the oscillation of the electrons around an equilibrium position. The electron is then subject to the elastic force FL described by: FL = −Kx

(1.3)

FC = −eE = −eE0e–iwt

(1.4)

where K is the elastic constant and x is the electron position. As the electrons have an electric charge −e, with e = 1.602 × 10−19 C, they are also subject to an oscillating electric field when they are excited by an electromagnetic wave and thus to a force FC: where E = E0(w)e–iwt is the excitation electric field, E0 is its amplitude, and w is the frequency of the oscillating electric field. The electron motion can be calculated using the fundamental principle of dynamics, as follows:

mẍ = FD + FL + FC = −

m x − Kx − eE0e–iwt (1.5) t

3

4

Introduction to Plasmonics

1 K x + x = -eE0e - iwt t m This differential equation can be solved easily, and the electron motion can be described as:  x+

or

x(t) = x0(w) e–iwt and x0(w)=

eE0

wˆ Ê m Á w 2 - w 02 + i ˜ Ë t¯

(1.6)

K , the eigen frequency. m Considering this solution, the electrons will be submitted to an oscillation. The frequency of oscillation w will be the same as that of the excitation electric field. Moreover, the oscillation amplitude x0 depends on the excitation frequency and the material characteristics (K, m, t). This oscillation amplitude is maximum at the w0 pulsation, which means that the absorption of light will be maximum and will reach a resonance condition.  The electron motion induces a dipolar moment P in the material, proportional to the electron motion, x(t), and given by the direction  of the electrical field E :  x   P = −ne e x(t) u = −ne e 0 E (1.7) E0 where e is the electron charge, ne is the electron density inside the  material, and u is the unitary vector in the motion direction.  P can also be written as:    P = e 0 c e E = e 0 Na E (1.8) where w 02 =

where e0 is the dielectric constant in the vacuum (e0= 8.85418782 × 10−12A2s4 kg−1m−3), ce is the electrical susceptibility, N is the number of dipoles, and a is the polarizability. The electric induction in the material is known as:      D = e 0 E + P = e 0 E + e 0 Na E (1.9)    D = e0e r E = e0 (1 + c e )E (1.10)

where er is the relative dielectric constant of the material. One can then deduce the following from Eqs. (1.6), (1.7), (1.9), and (1.10):

er = 1 + ce = 1 + ne ea = 1 −

ne ex0 (w ) e 0 E0

(1.11)

Surface Plasmons



er = 1 −

w p2

w t is the plasma frequency. w 2 - w 02 + i

(1.12)

ne e2 me0 In the ideal case, the free electron gas is without damping and has an w0 equal to 0 as described by the Drude model. The relative dielectric constant can be approximated by: where w p2 =



er = 1 −

w p2

w p2

= 1− 2 (1.13) w w t When er > 0, the electrons are displaced in the opposite direction of E. The field polarizes the material that acts as a dielectric. When er < 0, the electrons are displaced in the direction of E and create a depolarized field in the material that acts as a metal. w2 + i

1.2 Volume Plasmons

In 1920, Rudberg excited metals with an electron beam and measured the energies of the electrons after their interaction with the metal [7]. He noticed losses in electron energy at specific energy levels. The discretization of the energy lost by the free electrons of the metal reflects the quantification of the plasma oscillation energy. The quantum of the collective excitation of electron gases was then called plasmon, and its energy is a multiple of hwp. In the bulk metal, the electrons then oscillate, collectively forming a plasmon. This plasmon is infinitely expanded in the metal as the entire electron cloud oscillates at the plasma frequency wp in the metal volume.

1.3 Surface Plasmons

1.3.1 Plasmon Dispersion Curve Surface plasmon can also be excited at the interface between a dielectric medium and a metal, and an electromagnetic wave is formed, which can propagate at the surface of the metal. Such a plasmon is known as delocalized surface plasmon (DSP) or surface

5

6

Introduction to Plasmonics

plasmon polariton (SPP) [1, 2, 8]. The SPP can be excited using a transverse magnetic (TM) polarized electromagnetic wave. The k-vector of the excitation light has also to be equal the SPP k-vector to reach the phase matching between both k-vectors. Let us consider an electromagnetic wave at a dielectric/metal interface as presented in Fig. 1.1 with em and ed, the dielectric constants of the metal and the dielectric medium, respectively. The two electric fields in the dielectric and metallic media (Ed and Em, respectively) can be written as: Ed = (Exd, 0, Ezd) ei (kxd x + kzd z -wt ) (1.14) Em = (Exm, 0, Ezm) e (

i kxm x + kzm z -w t )

(1.15)

Figure 1.1 A schematic representing the excitation of a delocalized surface plasmon at the interface between a metal and a dielectric. em and ed are the dielectric constants of the metal and the dielectric medium, respectively.

Using Maxwell’s equations and the field continuity relations at the interface, the fields can be expressed as: Exd = Exm and Hxd = Hxm

Dzd = edEzd = em Ezm = Dzm

(1.16) (1.17)

Surface Plasmons

kzd kzm + = 0 (1.18) em ed One can notice a discontinuity of the component of the electric field along the z-direction (perpendicular to the surface). This discontinuity will be at the origin of the plasmon excitation, and this explains why we need to have a TM-polarized electromagnetic wave to create the plasmon. The dispersion relation of light can be expressed in the dielectric or metal media as:

2

Êwˆ 2 2 kxm + kzm = e m Á ˜ (1.19) Ëc¯



2



Êwˆ 2 2 kxd + kzd = e d Á ˜ (1.20) Ëc¯



Êe ˆ 2 2 kzm = Á m ˜ kzd (1.21) Ë ed ¯

Assuming that the k-vectors are equal in the x direction (kxd = kxm = kx = kSP), and that, from Eq. (1.18): 2

one can deduce that:

2 kzm

2

2 Ê e ˆ 2 Ê em ˆ Êwˆ = e m Á ˜ - k x2 = Á m ˜ kzd =Á Ëc¯ Ë ed ¯ Ë e d ˜¯

2

and that the dispersion relation of the SPP is:

È Ê w ˆ2 ˘ Íe d Á ˜ - k x2 ˙ (1.22) ÍÎ Ë c ¯ ˙˚

em + ed (1.23) emed

w = c kx

where c is the light velocity in the vacuum and w is the SPP frequency. Referring to the Drude model and replacing em with the er calculated previously (Eq. 1.13): Ê w = 1 Á e dw p2 + (1 + e d ) c 2k x2 - ÊÁ e dw p2 + (1 + e d ) c 2k x2 Ë 2e d Á Ë

(

1 1ˆ2 2 ˆ2 - 4e dw p2c 2k x2 ˜ ˜

)

¯ ˜ ¯

(1.24)

7

8

Introduction to Plasmonics

For low values of the wave vector kx, the plasmon dispersion curve increases nearly linearly, whereas for high values of kx, the curve reaches a plateau and becomes tangential to the plasma frequency wp/ 1 + e d .

1.3.2 Surface Plasmon Characteristics

As the dielectric constant of metal is negative (em < 0), the plasmon frequency is always lower than the light frequency in the dielectric medium or in the metal:

w Plasmon = ck x

em + ed 1 < ck x = w Light (1.25) e med e m or d

The dispersion curve of the surface plasmon is then always below the dispersion curve of the light (Fig. 1.2).

ω

kx Figure 1.2 Dispersion curve of the surface plasmon polariton at the interface between the metal and the dielectric medium.

If we consider the wave vector in the z-direction for both materials (dielectric medium and metal), we find from Eqs. (1.19), (1.20), and (1.25) that: 2



Êwˆ 2 2 kzm or zd = e m or d Á ˜ - k x < 0 (1.26) Ëc¯

Surface Plasmons

The wave vector in the z-direction is then imaginary in both dielectric and metal media. This means that the surface plasmon is an evanescent wave and that its amplitude decreases exponentially in the direction perpendicular to the metallic surface. The plasmon is then confined at the dielectric–metal interface. The decay length of the plasmon perpendicular to the surface is between 10 and 50 nm in the metal and about 100–200 nm in the dielectric media. However, the SPP can propagate along the surface at a distance of several tens of micrometers depending on the damping in the metal. The propagative length, LSP, is determined by the imaginary part of the wave vector in the surface direction such as:

LSP =

1 (1.27) 2Im (k x )

For instance, the values of LSP are round 22 µm and 500 µm for silver at 514 and 1060 nm, respectively. The plasmon is then delocalized in the plane of the interface between the dielectric medium and the metal but is confined (evanescent wave) in the direction perpendicular to the interface.

1.3.3 Surface Plasmon Excitation

To excite delocalized surface plasmons, both frequencies and wave vectors of the light and the plasmon have to be equal (phase matching) and thus the light dispersion curve w = ck has to intercept the SPP dispersion curve. However, the light dispersion curve is always above the SPP one (Fig. 1.2). To reach the curve intersection, the light wave vector value has to be increased.

1.3.3.1 Plasmon excitation using a grating

One way to increase the wave vector is to include a grating with a grating constant of p on the metal surface. The grazing orders of diffraction of the grating will be added to the excitation wave vector kxlight. The wave vector at the interface becomes: 2p (1.26) p where q is the incidence angle of the excitation light and m is the order of diffraction. The value of kx can then be increased by the



k x = klight sinq ± m

9

10

Introduction to Plasmonics

diffraction order wave vector. It will induce a shift in the light curve, which can then intercept the plasmon dispersion curve to excite the plasmon [9].

1.3.3.2 Plasmon excitation using a prism

Another way is to increase the refractive index of the dielectric medium at the interface with the metal. Otto [10] and Kretschmann [11] proposed two different configurations to realize the light/SPP phase matching by using a prism close to the metal surface. The Otto configuration was the first one to be proposed to excite the SPP. It uses a metallic film placed close but not in contact with the prism interface (Fig. 1.3a) forming an air layer between the prism and the metal surfaces. The excitation light goes through the prism with an angle to be in total internal reflection. An evanescent wave is created in the air at the prism surface. This evanescent wave will then be used to excite the SPP at the metal–air interface (Fig. 1.3a). This configuration is difficult to realize as it is hard to control a thin and homogeneous distance between the prism and the metal.

Figure 1.3 Schemes of (a) the Otto configuration and (b) the Kretschmann configuration.

For the Kretschmann configuration, a thin metallic film is deposited directly on the prism. As for the Otto configuration, the excitation light goes through the prism with an angle of total reflection at the prism–metal interface. An evanescent wave is created through the metal film that excites the metal in the opposite side of the film at the air–metal interface (Fig. 1.3b). The Kretschmann configuration is easier to realize than the Otto configuration, and thus it is the one currently used to excite SPP. With such Kretschmann configuration, the surface plasmon can be determined easily. A laser beam is used to illuminate the metal

Localized Surface Plasmon

film with a fixed excitation wavelength, and the reflected intensity is measured. The wave vector at the interface between the metal and the prism is defined as: kx = nprism klight sinq

(1.27)

where q is the incidence angle of the excitation light, klight is the incident light wavevector, and nprism is the optical index of the prism. As klight is fixed by the laser wavelength and nprism is constant, kx can be modified by changing the incident angle q. By varying q, one can reach the phase matching between the excitation light and the plasmon. In this case, one can observe a drastic decrease in the reflected intensity as the light energy is absorbed by the metal and transferred to the plasmon process. This intensity decrease is known as surface plasmon resonance (SPR). For instance, a gold thin film with a thickness of 50 nm and deposited on a glass prism (n = 1.5) exhibits an SPR for an incident angle close to 70° with an excitation wavelength of 660 nm. From Eq. (1.23), one can notice that the plasmon dispersion curve depends on the dielectric constant of the dielectric medium above the gold film. Thus, if we modify this dielectric constant through the absorption or the grafting of molecules at the surface of the film, we will observe a shift in the dispersion curve and as a consequence of the SPR. This concept is at the basis of the SPR-sensing method [12, 13]. This method is essentially used to measure the affinity constant between two molecules: One is grafted at the gold surface, whereas the other is injected at different concentrations to characterize the association or dissociation constants of the molecular interaction. The local modification of the dielectric constant can be quantified by measuring the angular shift or the variation in the reflectivity intensity and correlated with the number of molecules deposited at the gold surface. This method can be qualitative (interaction or not) and quantitative (number of interacting molecules) and is largely used in biology to observe antibody/antigen interactions.

1.4 Localized Surface Plasmon

When a nanostructure such as a nanoparticle is excited by an electromagnetic wave, the penetration depth of the electromagnetic field is equal to or higher than the nanoparticle size. The plasmon

11

12

Introduction to Plasmonics

can no longer propagate over large distances, as in the case of the delocalized surface plasmon. It is then confined inside the nanostructure volume and becomes a localized oscillation. In this case, it is known as localized surface plasmon (LSP) and can be observed on small structures that can trap the surface plasmons, such as nanocylinders, nanospheres, nanorods, nanoholes, or even topographical singularities on surfaces [14]. Thus, the LSP corresponds to an eigen mode of oscillation of the free electrons inside the nanostructures and is confined in their three dimensions. When the light frequency is equal to the eigen mode frequency, a resonance effect occurs and the LSP is excited. The position of the LSP resonance (LSPR) depends not only on the nature of the metal and the direct environment of the nanostructure (i.e., the surrounding medium), but also on the size (Fig. 1.4), shape, and spatial distribution of metallic nanostructures (coupled with other nanostructures). For instance, Fig. 1.4 shows that the plasmon resonance of nanocylinders is redshifted when the nanocylinder diameter increases from 100 to 200 nm [15].

Figure 1.4 Extinction spectra of gold nanocylinders on glass surface. Diameters varying between 100 nm and 200 nm, height equals to 50 nm, and the gap between two nanocylinders is 200 nm in both directions (extracted from Ref. [15]).

The LSPR can be measured by extinction spectroscopy. The extinction spectrum corresponds to the intensity attenuation experienced by a light beam after interaction with a nanostructure. The extinction process results in two kinds of light interaction with the nanostructures: the absorption and the scattering of the

Localized Surface Plasmon

incident light by the material. These two processes are also known as nonradiative and radiative interactions, respectively, as photons may or may not be emitted during the process. The extinction cross section sext is then defined as:

sext = sabs + sscat

(1.27)

where sabs and sscat are the absorption and scattering cross sections of the material embedded in a dielectric medium. When plasmon is excited, the extinction spectrum exhibits a large band (width of a few tens of nanometers). The maximum of this band corresponds to the largest interaction of light with the nanoparticles and thus to the plasmon resonance. In the following, we will focus on the plasmon resonance of different kinds of nanostructures (nanospheres, elongated nanoparticles) and also on the influence of the different geometrical (size and shape), chemical (nature of the metal), and environmental (surrounding medium, coupling) parameters on the position of the plasmon resonance.

1.4.1 LSP in Spherical Nanoparticles

The electromagnetic wave interaction with the spherical particles can be completely determined using the Mie theory [16−20]. It solves analytically Maxwell’s equation in the spherical coordinate system to calculate the extinction and scattering cross sections (the absorption one is then determined by sabs = sext − sscat). Mie also introduced the extinction, absorption, and scattering efficiencies, 𝑄ext, 𝑄abs, and 𝑄scat, of the spherical nanoparticles defined as:

𝑄 = s/p𝑅2

(1.27)

and that corresponds to the cross section normalized by the intersecting surface of the nanosphere (radius R). The Mie theory allows to determine the position and the width of the plasmon resonance as well as the contribution of both absorption and scattering depending on the size, the chemical nature of the particle, and the surrounding medium nature. Small nanoparticles give rise to the distribution of dipolar charges on the surface. In the case of spherical particles, it presents one resonance mode due to homogeneity in charge displacement. The symmetry gives a resonance independent of the polarization direction of the incident light and only one plasmon band is observable.

13

Figure 1.5 Extinction, absorption, and scattering efficiencies versus the excitation wavelength for gold nanospheres in vacuum with different diameters.

14 Introduction to Plasmonics

Localized Surface Plasmon

As shown in Fig. 1.5, the nanosphere size plays a role in the resonance response. First, the position of the resonance depends on the size and an increase in particle size will induce a redshift of the resonance. Second, the contribution of the absorption and the scattering to the extinction process will also depend on the size (Fig. 1.5 and Fig. 1.6). For a particle with a radius smaller than 40 nm, absorption dominates while scattering is negligible. The energy absorption in small nanoparticles is first due to the surface plasmons mode (band between 350 and 400 nm) and second due to the electron interband transitions (band lower than 300 nm). On the contrary, in the case of nanoparticles with a bigger size, scattering dominates the resonance response whereas absorption becomes negligible [21].

Figure 1.6 (a) Absorption and (b) scattering efficiencies versus the excitation wavelength for silver nanospheres in vacuum with radius from 5 nm to 50 nm [21].

Thus, the Mie model provides a complete description of the light– matter interaction with small particles. However, it can be tricky to use as it is necessary to calculate the whole set of Mie coefficients, which could be difficult to interpret straightly. Some simpler models can be used to approximate the nanosphere behavior and to have a better insight into the different processes that occur during the plasmon excitation as well as into the influence of the different parameters. The first approximation that can be used is the Rayleigh or quasistatic approximation. In this way, the nanoparticle size is considered negligible compared to the wavelength (R 420 nm)

R2

R2 R1

Scheme 3.8 The Suzuki–Miyaura reaction. Adapted from Ref. [88].

Zhu et al. [89] studied the influence of a plasmonic anisotropic nanocatalyst that consisted of a gold nanopyramid core covered at the extremities with silica and on the sides with palladium. This nanocatalyst was employed in the Suzuki–Miyaura cross-coupling, and the reaction was conducted in water at 50°C. After irradiation with an 808 nm laser for 1 h, the conversion of the reaction was around 85% versus 65% under dark conditions (Scheme 3.9). Br +

OH B OH

Au NBPs/sidePd-endSiO2 K2CO3 (3 eq), H2O , 1 h, 50 °C Laser 3 W - 808 nm

Scheme 3.9 The Suzuki–Miyaura reaction with Au nanopyramids coated with silica and palladium. Adapted from Ref. [89].

73

74

Surface Plasmon Resonance‒Mediated Organic Synthesis

Recently, Pd nanoflowers were synthesized via the reduction of Pd(II)(acac)2 in ethanol under a CO atmosphere [90]. These nanoflowers displayed a broad absorption plasmon band in the visible and near-IR ranges, and their plasmonic properties under visible-light irradiation in Suzuki–Miyaura cross-coupling reactions were studied. The reaction was carried out using iodobenzene, phenylboronic acid, and Cs2CO3 as the base under visible-light irradiation, and the results were compared with the reaction conducted in the dark. Under light irradiation (300 W Xe lamp equipped with a water cell filter to absorb the near-IR radiation, and a 475 nm cutoff filter to avoid direct photoactivation of the aromatic compounds), the conversion was 96% versus 53% in the dark after 240 min. The authors explained that the plasmonic effect increases the reactivity of the palladium nanostructures, which induces heat and hot electrons on the Pd surface. Koohgard et al. [91] studied PdNPs deposited onto TiO2 through a photodeposition method under solar light that exhibited plasmonic properties. In aqueous media (water:PEG = 1:1), 4-methoxyiodobenzene reacted with phenylboronic acid to give the final product with 91% yield under visible-light irradiation (400–750 nm, LED, 15 W, 28°C) compared to 22% yield in dark conditions after 4 h. Photogenerated electron– hole pairs formed under visible-light irradiation are the main parameters of activation of the different substrates in the Suzuki– Miyaura reaction. Gangishetty et al. [92] used AuPd bimetallic nanotriangles, and under green LED irradiation, the Suzuki–Miyaura reaction rate was improved compared to the dark reaction after only a few hours, and this enhancement was mainly due to plasmonic heating effects, with a minor contribution from plasmonic hot electrons. Nemygina et al. [93] synthetized a nanocatalyst composed from 20 nm core AuNPs surrounded by a shell of 1 nm palladium NPs and immobilized within hyper cross-linked polystyrene. Under visible-light irradiation using a 300 W filament lamp, the product yield of the Suzuki–Miyaura coupling was 1.3 times higher than in the dark (from 55.3% to 71.6%). Furthermore, Rohani et al. [94] designed a hydrogenated urchin-like yolk@shell TiO2 structure decorated with Au and Pd nanoparticles (HUY@S-TOH@AuPd), which was used in the Suzuki–Miyaura reaction. Under 300 W xenon lamp irradiation, 4-methyliodobenzene reacted with phenylboronic

Miscellaneous Chemical Reactions

acid in the presence of this nanocatalytic system with a 94% yield after 1 h, while under dark conditions, no product was formed. Very recently, Li et al. [95] focused on TERS to study the Suzuki–Miyaura reaction. In that regard, a monolayer of one reactant was deposited on the surface of Au@Pd NPs and then the second reactant was deposited on the scanning probe. Then the probe is brought close to the sample surface, which is illuminated continuously by a laser, and the reaction takes place and is monitored by Raman spectroscopy. Thus, they could probe within seconds the reactivity of different aryl halides under irradiation. This TERS technique allows fast, robust, and reliable analysis and understanding of molecular reactivity for plasmon-enhanced reactions. Finally, AuPd bimetallic nanowheels used in the oxidation of benzyl alcohol were proven to be efficient catalysts in the Suzuki coupling reaction [16]. Li also reported the encapsulation of core–shell Cu@Ni NPs into the zeolitic imidazolate framework ZIF-8 for the photocatalytic Suzuki coupling reaction [96]. This is one of the few examples of the incorporation of nanoparticles into metal–organic frameworks and allows the combination of high porosity, shape-selectivity, and catalytic enhancement. It is important to note that the C–B bond cleavage deboronation is a side reaction that can lower the Suzuki coupling reaction yield. Very recently, a protodeboronation reaction [97] was achieved using AuNPs and planar gold substrates under 785 nm light irradiation. The model reaction studied was the protodeboronation reaction from 4-mercaptophenylboronic acid (MPBA) to benzenethiol (BT) (Scheme 3.10), and this reaction occurs upon plasmon excitation in nanogaps between the Au substrate where a self-assembled monolayer (SAM) of MPBA was deposited followed by AuNP deposition on the SAM. A schematic representation of the setup is presented in Fig. 3.17. HO

B

OH

H

Nanogap hν SH

SH

Scheme 3.10 Protodeboronation of MPBA to BT in nanogaps upon plasmon excitation.

75

76

Surface Plasmon Resonance‒Mediated Organic Synthesis

Figure 3.17 Schematic of the experimental setup for the protodeboronation reaction. The sample is prepared by drop-casting AuNPs (51 nm) onto MPBA SAMs on 150 nm thick Au substrates. The sample is placed under a 50× objective in an upright microscope. A laser (785, 638, or 532 nm) is focused on the sample through the objective for plasmon excitation. Adapted from Ref. [97].

Among the different plausible mechanisms upon irradiation, the reaction by plasmonic heating was ruled out since no reaction took place upon heating to 100°C (without laser excitation). The reaction was more likely due to hot electrons and hot holes that directly transfer to MPBA, which subsequently leads to the formation of BT. Additionally, the authors observed that the presence of oxygen molecules promoted the reaction by opening another electrontransfer channel.

3.5.2 Photoconversion of CO2

Reduced graphene-coated AuNPs (r-GO AuNPs) were proven to be excellent photocatalysts for the photoconversion of CO2 into formic acid (HCOOH) (>90%) [98]. It was suggested that reduced graphene layers can play a role as efficient electron acceptors and transporters, facilitating the utilization of hot electrons for plasmonic photocatalysts. The femtosecond transient spectroscopic analysis also showed 8.7 times higher transport efficiency of hot plasmonic electrons in r-GO AuNPs compared with AuNPs.

Miscellaneous Chemical Reactions

Very recently, Li et al. [99] developed a Bi2O3−x catalyst that reduced CO2 to CO under infrared light. They used a low-intensity LED light at 940 nm and was able to reach a catalytic activity of 4.6 µmol/gh. It was also demonstrated that the photocatalytic activity of Bi2O3−x was due to a linear dependence on light intensity and temperature. Zhang et al. [100] reported the reduction of CO2 to CO using molecular hydrogen as the reducing agent and catalyzed by an AgZrO2 photocatalyst under UV-visible light. The reaction was studied mechanistically and spectroscopically and found to involve the dual roles of light: charge separation and plasmon heating.

3.5.3 Other Chemical Reactions

Combining an aldehyde, an amine, and phenylacetylene at a high temperature led to the formation of propargylamines in good yields (50–95%) using the excitation of the surface plasmons of AuNPs supported on oxide zinc (AuNP@ZnO) (Scheme 3.11) [101]. The products were obtained after 2 h under LED excitation at 530 nm. R1 N H

+

H

R2

+

O

hν (530 nm) AuNPs@ZnO

R2

R1 = CH2, C2H4, OCH2 R 2 = H, C6H5, CH(CH3)2

N R1

Scheme 3.11 Propargylamine synthesis catalyzed by AuNP@ZnO.

Hydroamination of alkynes or alkenes is considered an atomeconomical route to various organic nitrogen molecules and leads to fine chemicals and synthetic intermediates. The intermolecular reaction is much more challenging because it usually requires higher temperatures (generally above 100°C) than the intramolecular version to achieve high yields. However, at high temperatures, the reaction equilibrium shifts toward lower conversions, which creates a dilemma. Thus, Zhao et al. [102] studied the photocatalytic hydroamination of alkynes with aniline molecules using AuNPs supported on oxide nitrogen-doped titanium, AuNP@TiO2-N, which proved to be highly efficient and selective for the reaction at room

77

78

Surface Plasmon Resonance‒Mediated Organic Synthesis

temperature under green light irradiation (Scheme 3.12). In this reaction, both AuNPs and the support contribute to the activation of reactants: Aniline molecules preferentially interact with AuNP, while alkynes were adsorbed onto active support sites [102]. R1

H

+

R 2 NH2

AuNP@TiO 2-N hν

R2 R1

N

CH3

Scheme 3.12 Photocatalytic hydroamination of alkynes catalyzed by AuNP@ TiO2-N.

The combination of plasmonics and solvent-free chemical synthesis is an attractive green chemistry approach. Meng et al. reported the solvent-free hydroamination of alkynes using plasmonic AuNPs coupled with a TiO2 2D photonic layer on nanotube arrays (Fig. 3.18) [103]. The solvent-free hydroamination of phenylacetylene using aniline was performed with both Au/ TiO2 photonic crystal nanotubes (PCNT) and Au/TiO2 nanotubes. An absence of hydroamination was noted when only the TiO2 PCNT support was present, confirming that AuNPs are responsible for catalytic activity. The authors suggest that the simultaneous adsorption or activation of the reactants is required for the reaction to proceed with high turnover.

Figure 3.18 Hydroamination of phenylacetylene using AuNP-photonic crystal nanotubes. Adapted from Ref. [103].

The synthesis of 4-benzoylmorpholine from benzaldehyde and morpholine via an amide formation was catalyzed by AuNPs supported on SiO2 (AuNP@SiO2) at room temperature and under laser irradiation, and quantitative yields were obtained [104]. This reaction was also extended to a tandem oxidation/amidation process starting with benzyl alcohol and morpholine. In this case, the catalyst irradiation at 532 nm leads to the initial formation of benzaldehyde, through oxidation by hydrogen peroxide present in the middle, and then the amidation reaction took place (Scheme 3.13).

Miscellaneous Chemical Reactions

O OH

HN +

AuNP@SiO 2 O

N O

hν, r.t.

Scheme 3.13 Laser-driven tandem oxidation/amidation reaction catalyzed by AuNP@SiO2.

This reaction was also used in the synthesis of methyl-Nbenzylbenzamide (from methyl benzylamine) and N-benzylamide (the benzylamine) with high performance and selectivity. A number of other catalysts possessing a plasmon band, such as gold nanorods and gold nanoparticles supported on various oxides, were also investigated; however, low yields were obtained. The highest performing catalyst was AuNP@SiO2, which preserved 90% of its initial activity after a minimum of two reuses. Bazyar et al. [105] have shown that Au@ZnO core−shell NPs were good photoredox catalysts for the facile trifluoromethylation of a wide range of substrates using sodium trifluoroacetate (Scheme 3.14). The method was extended to the trifluoromethylation of biologically active molecules, and one reaction was even performed at the 10 g scale [105]. For the mechanism, they proposed that gold absorbs visible light, and a hot electron is then transferred to the conduction band (CB) of ZnO, which prevents the electron/hole (e−/h+) recombination. This photogenerated h+ on gold can oxidize CF3CO2Na, generating the CF3 radical through decarboxylation. R

CF3

Au@ZnO, CF3CO2Na 11 W blue LED (λ = 400−495 nm) R = I, Br, H, B(OH)2 Scheme 3.14 Photocatalytic trifluoromethylation under visible light catalyzed by AuNP@ZnO.

The copper-catalyzed 1,3-dipolar azide–alkyne cycloaddition reaction, known as Huisgen reaction and commonly called “clickchemistry,” allows the formation of triazole molecules. The catalytically active species in this reaction is Cu(I), but Cu(I) is usually not stable when heated or exposed to oxygen. Thus, this

79

80

Surface Plasmon Resonance‒Mediated Organic Synthesis

reaction is usually conducted with Cu(II) species and a reducing agent. Sun et al. [106] studied this reaction using AuNPs and Cu(II) precursors under irradiation, and very good yields were obtained (Scheme 3.15). The observed catalytic enhancement of this reaction was due to the synergistic effect of LSPR inducing a local heating effect, and plasmonic hot electrons are formed, which reduce Cu(II) species to catalytically active Cu(I) species. Au–CuO hybrid nanostructures were described that proved to be an efficient means for coupling visible-light energy, which heterogeneously catalyzes the “click-chemistry” reaction, without the use of any conventional heating and a chemical reducing agent. N N N

N3 +

AuNP, CuSO4, visible light Or Au-CuO, visible light

Scheme 3.15 Surface plasmon resonances enhanced copper catalyzed 1,3-dipolar azide–alkyne cycloaddition reaction.

Very recently, Guselnikova et al. [107] studied the influence of temperature on plasmon-driven click-chemistry. The alkyne reagent was grafted on the surface of spherical AuNPs through a diazonium approach, and this nanotemplate was dispersed in acetonitrile and reacted with 4-azidobenzoic acid under light irradiation. When they performed the reaction at –35°C, they noticed an important increase in reactivity, which was unexpected. To explain this phenomenon, they performed DFT calculations and proposed that the decrease in electron–phonon scattering and retardation of organic molecule relaxation after plasmon-induced excitation induced this reversed dependence on the reaction rate. Thiol-ene coupling is part of the “click-chemistry” set of reactions and consists of the reaction of alkene with a thiol to form an alkyl sulfide. Tijunelyte et al. [108] studied this thiol-ene “click” reaction using gold nanocylinders under visible light. 1,4-butanedithiol, bearing a double-bond terminal group, was initially self-assembled onto the Au cylinders. The modified surface was then exposed to thiophenol in the liquid phase in the presence of a radical initiator (2,2′-azobis(2-methylpropionamidine)dihydrochloride, AAPH)

Miscellaneous Chemical Reactions

and irradiated at 660 nm (Scheme 3.16). Under irradiation, the reaction rate was accelerated up to 20 times and was interpreted as the combination of both photonic and thermal effects, which can enhance and control chemical reactions at the nanoscale level.

Scheme 3.16 Thio-ene click chemistry reaction taking place at the surface of laser-illuminated Au nanocylinders.

A Friedel–Crafts alkylation was performed using a nanocatalyst composed of AuNPs supported on either commercial niobium oxide HY 340 or mesoporous niobium oxide, these supports replacing the classical Lewis acid (Scheme 3.17) [109]. Under irradiation of this nanocatalyst, anisole and benzyl chloride reacted at 80°C and 100% conversion was obtained within 2 h. In the absence of LED irradiation, for the reaction to proceed, the temperature was increased beyond 120°C and the conversion reached 80% after 2 h, while at 80°C, no conversion was observed. O

+

Cl

Au NPs/Nb2O5

O

2 h, 80 °C green LED - 565 nm

Scheme 3.17 Friedel–Crafts reaction taking place at the surface of laserilluminated Au NPs/Nb2O5.

An interesting tandem oxidative coupling of alcohols and amines using Au/CeO2 nanorods promoted by visible light was reported by Teixeira et al. [110]. In this reaction, it is proposed that hot electrons generated from the LSPR on AuNPs are trapped by oxygen vacancies in the CeO2 nanorods, which are then picked up by O2 molecules on the surface. Concomitantly, hot holes generated by the AuNPs

81

82

Surface Plasmon Resonance‒Mediated Organic Synthesis

promote a-hydride abstraction from the alcohol, yielding a nascent aldehyde, which in turn couples with an amine (Fig. 3.19).

Figure 3.19 Proposed mechanism for Au–CeO2-mediated coupling of benzyl alcohol with aniline under visible-light irradiation. Adapted from Ref. [110].

Figure 3.20 Proposed mechanism for the photocatalytic oxidative crossesterification of methanol and benzyl alcohol. Adapted from Ref. [111].

References

Alcohols were also found to react with methanol in an LSPRpromoted oxidative esterification reaction reported by Wang et al. [111]. In this chapter, layered double hydroxide Au–Cu NPs catalyzed the esterification of methanol with benzyl alcohol in the presence of molecular O2. Apparently, the presence of copper in the AuNPs promotes the formation of the superoxide radical O2•−, which is invoked in the mechanism shown in Fig. 3.20.

References

1. Gelle, A., Jin, T., de la Garza, L., Price, G. D., Besteiro, L. V., Moores, A. Applications of plasmon-enhanced nanocatalysis to organic transformations. Chem. Rev. 2020, 120(2), 986–1041. https://doi. org/10.1021/acs.chemrev.9b00187. 2. Gellé, A. and Moores, A. Plasmonic nanoparticles: Photocatalysts with a bright future. Curr. Opin. Green Sustain. Chem. 2019, 15, 60–66. https:// doi.org/10.1016/j.cogsc.2018.10.002.

3. Tian, Y. and Tatsuma, T. Mechanisms and applications of plasmoninduced charge separation at TiO2 films loaded with gold nanoparticles. J. Am. Chem. Soc. 2005, 127(20), 7632–7637. https://doi.org/10.1021/ ja042192u. 4. Tanaka, A., Hashimoto, K., and Kominami, H. Selective photocatalytic oxidation of aromatic alcohols to aldehydes in an aqueous suspension of gold nanoparticles supported on cerium(IV) oxide under irradiation of green light. Chem. Commun. 2011, 47(37), 10446–10448. https:// doi.org/10.1039/C1CC13801D.

5. Tanaka, A., Hashimoto, K., and Kominami, H. Preparation of Au/CeO2 exhibiting strong surface plasmon resonance effective for selective or chemoselective oxidation of alcohols to aldehydes or ketones in aqueous suspensions under irradiation by green light. J. Am. Chem. Soc. 2012, 134(35), 14526–14533. https://doi.org/10.1021/ja305225s.

6. Tsukamoto, D., Shiraishi, Y., Sugano, Y., Ichikawa, S., Tanaka, S., and Hirai, T. Gold nanoparticles located at the interface of anatase/rutile TiO2 particles as active plasmonic photocatalysts for aerobic oxidation. J. Am. Chem. Soc. 2012, 134(14), 6309–6315. https://doi.org/10.1021/ ja2120647. 7. Zhai, W., Xue, S., Zhu, A., Luo, Y., and Tian, Y. Plasmon-driven selective oxidation of aromatic alcohols to aldehydes in water with recyclable Pt/TiO2 nanocomposites. ChemCatChem 2011, 3(1), 127–130. https:// doi.org/10.1002/cctc.201000303.

83

84

Surface Plasmon Resonance‒Mediated Organic Synthesis

8. Hallett-Tapley, G. L., Silvero, M. J., Gonzalez-Bejar, M., Grenier, M., NettoFerreira, J. C., and Scaiano, J. C. Plasmon-mediated catalytic oxidation of sec-phenethyl and benzyl alcohols. J. Phys. Chem. C 2011, 115(21), 10784–10790. https://doi.org/10.1021/jp202769a.

9. Hallett-Tapley, G. L., Silvero, M. J., Bueno-Alejo, C. J., González-Béjar, M., McTiernan, C. D., Grenier, M., Netto-Ferreira, J. C., and Scaiano, J. C. Supported gold nanoparticles as efficient catalysts in the solventless plasmon mediated oxidation of sec-phenethyl and benzyl alcohol. J. Phys. Chem. C 2013, 117(23), 12279–12288. https://doi.org/10.1021/ jp311069v. 10. Wee, T.-L. E., Schmidt, L. C., and Scaiano, J. C. Photooxidation of 9-anthraldehyde catalyzed by gold nanoparticles: Solution and single nanoparticle studies using fluorescence lifetime imaging. J. Phys. Chem. C 2012, 116(45), 24373–24379. https://doi.org/10.1021/jp308956y.

11. Sarina, S., Zhu, H., Jaatinen, E., Xiao, Q., Liu, H., Jia, J., Chen, C., and Zhao, J. Enhancing catalytic performance of palladium in gold and palladium alloy nanoparticles for organic synthesis reactions through visible light irradiation at ambient temperatures. J. Am. Chem. Soc. 2013, 135(15), 5793–5801. https://doi.org/10.1021/ja400527a.

12. Eerdemutu, Ding, L., Li, C., Zhao, S., and Sheng, X. Synthesis of imines by selective oxidation of amines on alloy nanoparticles of palladium and gold under visible light irradiation at ambient temperatures. Catal. Lett. 2020, 150(6), 1757–1765. https://doi.org/10.1007/s10562-01903079-w. 13. Swaminathan, S., Rao, V. G., Bera, J. K., and Chandra, M. The pivotal role of hot carriers in plasmonic catalysis of C−N bond forming reaction of amines. Angew. Chem. Int. Ed. 2021, 60(22), 12532–12538. https://doi. org/10.1002/anie.202101639.

14. Zhang, K., Xi, Z., Wu, Z., Lu, G., and Huang, X. Visible-light-induced selective oxidation of amines into imines over UiO-66-NH2@Au@COF core–shell photocatalysts. ACS Sustain. Chem. Eng. 2021, 9(37), 12623– 12633. https://doi.org/10.1021/acssuschemeng.1c04112.

15. Kumar, A., Kumari, N., Dubbu, S., Kumar, S., Kwon, T., Koo, J. H., Lim, J., Kim, I., Cho, Y.-K., Rho, J., and Lee, I. S. Nanocatalosomes as plasmonic bilayer shells with interlayer catalytic nanospaces for solar-lightinduced reactions. Angew. Chem. Int. Ed. 2020, 59(24), 9460–9469. https://doi.org/10.1002/anie.202001531. 16. Huang, X., Li, Y., Chen, Y., Zhou, H., Duan, X., and Huang, Y. Plasmonic and catalytic AuPd nanowheels for the efficient conversion of light into chemical energy. Angew. Chem. Int. Ed. 2013, 52(23), 6063–6067. https://doi.org/10.1002/anie.201301096.

References

17. Jiang, T., Jia, C., Zhang, L., He, S., Sang, Y., Li, H., Li, Y., Xu, X., and Liu, H. Gold and gold–palladium alloy nanoparticles on heterostructured TiO2 nanobelts as plasmonic photocatalysts for benzyl alcohol oxidation. Nanoscale 2014, 7(1), 209–217. https://doi.org/10.1039/ C4NR05905K. 18. Li, H., Qin, F., Yang, Z., Cui, X., Wang, J., and Zhang, L. New reaction pathway induced by plasmon for selective benzyl alcohol oxidation on BiOCl possessing oxygen vacancies. J. Am. Chem. Soc. 2017, 139(9), 3513–3521. https://doi.org/10.1021/jacs.6b12850.

19. Liu, L., Ouyang, S., and Ye, J. Gold-nanorod-photosensitized titanium dioxide with wide-range visible-light harvesting based on localized surface plasmon resonance. Angew. Chem. Int. Ed. 2013, 52(26), 6689– 6693. https://doi.org/10.1002/anie.201300239. 20. Nadeem, M. A. and Idriss, H. Photo-thermal reactions of ethanol over Ag/TiO2 catalysts. The role of silver plasmon resonance in the reaction kinetics. Chem. Commun. 2018, 54(41), 5197–5200. https://doi. org/10.1039/C8CC01814F.

21. Kowalska, E., Mahaney, O. O. P., Abe, R., and Ohtani, B. Visible-lightinduced photocatalysis through surface plasmon excitation of gold on titania surfaces. Phys. Chem. Chem. Phys. 2010, 12(10), 2344–2355. https://doi.org/10.1039/B917399D. 22. Kowalska, E., Abe, R., and Ohtani, B. Visible light-induced photocatalytic reaction of gold-modified titanium(IV) oxide particles: Action spectrum analysis. Chem. Commun. 2009, 2, 241–243. https://doi.org/10.1039/ B815679D.

23. Sugano, Y., Shiraishi, Y., Tsukamoto, D., Ichikawa, S., Tanaka, S., and Hirai, T. Supported Au–Cu bimetallic alloy nanoparticles: An aerobic oxidation catalyst with regenerable activity by visible-light irradiation. Angew. Chem. Int. Ed. 2013, 52(20), 5295–5299. https://doi.org/10.1002/ anie.201301669. 24. Colmenares, J. C., Lisowski, P., Łomot, D., Chernyayeva, O., and Lisovytskiy, D. Sonophotodeposition of bimetallic photocatalysts Pd–Au/TiO2: Application to selective oxidation of methanol to methyl formate. ChemSusChem 2015, 8(10), 1676–1685. https://doi.org/10.1002/ cssc.201403125.

25. Sun, H., Chen, S., Yang, W., Wang, L., Tang, R., Zhang, X., Zheng, R., Gu, S., Jiang, Y., Liang, W., and Huang, J. Plasmon-enhanced alcohol oxidations over porous carbon nanosphere-supported palladium and gold bimetallic nanocatalyst. Appl. Catal. B Environ. 2021, 292, 120151. https://doi.org/10.1016/j.apcatb.2021.120151.

85

86

Surface Plasmon Resonance‒Mediated Organic Synthesis

26. Wang, Z., Du, J., Zhang, Y., Han, J., Huang, S., Hirata, A., and Chen, M. Free-standing nanoporous gold for direct plasmon enhanced electrooxidation of alcohol molecules. Nano Energy 2019, 56, 286–293. https://doi.org/10.1016/j.nanoen.2018.11.060. 27. Zhu, P., Sun, X., Wang, Y., Zhang, J., Gu, X., and Zheng, Z. Multifunctional oxygen vacancies in WO3–x for catalytic alkylation of C–H by alcohols under red-light. J. Catal. 2021, 402, 208–217. https://doi.org/10.1016/j. jcat.2021.08.040.

28. Abad, A., Corma, A., and García, H. Supported gold nanoparticles for aerobic, solventless oxidation of allylic alcohols. Pure Appl. Chem. 2007, 79(11), 1847–1854. https://doi.org/10.1351/pac200779111847. 29. Landry, M. J., Gellé, A., Meng, B. Y., Barrett, C. J., and Moores, A. Surfaceplasmon-mediated hydrogenation of carbonyls catalyzed by silver nanocubes under visible light. ACS Catal. 2017, 7(9), 6128–6133. https://doi.org/10.1021/acscatal.7b02128. 30. Christopher, P., Xin, H., and Linic, S. Visible-light-enhanced catalytic oxidation reactions on plasmonic silver nanostructures. Nat. Chem. 2011, 3(6), 467–472. https://doi.org/10.1038/nchem.1032.

31. Marimuthu, A., Zhang, J., and Linic, S. Tuning selectivity in propylene epoxidation by plasmon mediated photo-switching of cu oxidation state. Science 2013, 339(6127), 1590–1593. https://doi.org/10.1126/ science.1231631.

32. Zhang, X., Kumari, G., Heo, J., and Jain, P. K. In situ formation of catalytically active graphene in ethylene photo-epoxidation. Nat. Commun. 2018, 9(1), 3056. https://doi.org/10.1038/s41467-01805352-9. 33. Ide, Y., Matsuoka, M., and Ogawa, M. Efficient visible-light-induced photocatalytic activity on gold-nanoparticle-supported layered titanate. J. Am. Chem. Soc. 2010, 132(47), 16762–16764. https://doi. org/10.1021/ja1083514. 34. Zheng, Z., Huang, B., Qin, X., Zhang, X., Dai, Y., and Whangbo, M.-H. Facile in situ synthesis of visible-light plasmonic photocatalysts M@TiO2 (M = Au, Pt, Ag) and evaluation of their photocatalytic oxidation of benzene to phenol. J. Mater. Chem. 2011, 21(25), 9079–9087. https://doi. org/10.1039/C1JM10983A.

35. Xiao, Y., Liu, J., Mai, J., Pan, C., Cai, X., and Fang, Y. High-performance silver nanoparticles coupled with monolayer hydrated tungsten oxide nanosheets: The structural effects in photocatalytic oxidation of cyclohexane. J. Colloid Interface Sci. 2018, 516, 172–181. https://doi. org/10.1016/j.jcis.2018.01.057.

References

36. Takeuchi, Y., Fujita, T., and Takeyasu, N. Plasmon-mediated chemical transformation from alkane to alkene on a silver nanoparticle array under 532 Nm excitation. Phys. Chem. Chem. Phys. 2019, 21(14), 7502– 7507. https://doi.org/10.1039/C9CP00129H. 37. Lee, S. W., Lee, C., Goddeti, K. C., Kim, S. M., and Park, J. Y. Surface plasmon-driven catalytic reactions on a patterned Co3O4/Au inverse catalyst. RSC Adv. 2017, 7(88), 56073–56080. https://doi.org/10.1039/ C7RA10450B. 38. Lee, S. W., Hong, J. W., Lee, H., Wi, D. H., Kim, S. M., Han, S. W., and Park, J. Y. The surface plasmon-induced hot carrier effect on the catalytic activity of CO oxidation on a Cu2O/hexoctahedral Au inverse catalyst. Nanoscale 2018, 10(23), 10835–10843. https://doi.org/10.1039/C8NR00555A. 39. Zhang, Y., Li, Q., Liu, C., Shan, X., Chen, X., Dai, W., and Fu, X. The promoted effect of a metal-organic frameworks (ZIF-8) on Au/TiO2 for CO oxidation at room temperature both in dark and under visible light irradiation. Appl. Catal. B Environ. 2018, 224, 283–294. https://doi. org/10.1016/j.apcatb.2017.10.027.

40. Dong, B., Fang, Y., Xia, L., Xu, H., and Sun, M. Is 4-nitrobenzenethiol converted to p,p′-dimercaptoazobenzene or 4-aminothiophenol by surface photochemistry reaction? J. Raman Spectrosc. 2011, 42(6), 1205–1206. https://doi.org/10.1002/jrs.2937. 41. Dong, B., Fang, Y., Chen, X., Xu, H., and Sun, M. Substrate-, wavelength-, and time-dependent plasmon-assisted surface catalysis reaction of 4-nitrobenzenethiol dimerizing to p,p′-dimercaptoazobenzene on Au, Ag, and Cu films. Langmuir 2011, 27(17), 10677–10682. https://doi. org/10.1021/la2018538.

42. Ren, X., Cao, E., Lin, W., Song, Y., Liang, W., and Wang, J. Recent advances in surface plasmon-driven catalytic reactions. RSC Adv. 2017, 7(50), 31189–31203. https://doi.org/10.1039/C7RA05346K. 43. Zhao, L.-B., Zhang, M., Huang, Y.-F., Williams, C. T., Wu, D.-Y., Ren, B., and Tian, Z.-Q. Theoretical study of plasmon-enhanced surface catalytic coupling reactions of aromatic amines and nitro compounds. J. Phys. Chem. Lett. 2014, 5(7), 1259–1266. https://doi.org/10.1021/ jz5003346.

44. Zhu, H., Ke, X., Yang, X., Sarina, S., and Liu, H. Reduction of nitroaromatic compounds on supported gold nanoparticles by visible and ultraviolet light. Angew. Chem. Int. Ed. 2010, 49(50), 9657–9661. https://doi. org/10.1002/anie.201003908. 45. Liu, Z., Huang, Y., Xiao, Q., and Zhu, H. Selective reduction of nitroaromatics to azoxy compounds on supported Ag–Cu alloy nanoparticles through

87

88

Surface Plasmon Resonance‒Mediated Organic Synthesis

visible light irradiation. Green Chem. 2016, 18(3), 817–825. https:// doi.org/10.1039/C5GC01726B.

46. Liu, Q., Zhang, J., Xing, F., Cheng, C., Wu, Y., and Huang, C. Plasmonenhanced and controllable synthesis of azobenzene and hydrazobenzene using Au/TiO2 composite. Appl. Surf. Sci. 2020, 500, 144214. https:// doi.org/10.1016/j.apsusc.2019.144214.

47. Ding, Q., Chen, M., Fang, Y., Zhang, Z., and Sun, M. Plasmon-driven diazo coupling reactions of p-nitroaniline via −NH2 or −NO2 in atmosphere environment. J. Phys. Chem. C 2017, 121(9), 5225–5231. https://doi. org/10.1021/acs.jpcc.6b12589. 48. Ke, X., Zhang, X., Zhao, J., Sarina, S., Barry, J., and Zhu, H. Selective reductions using visible light photocatalysts of supported gold nanoparticles. Green Chem. 2012, 15(1), 236–244. https://doi. org/10.1039/C2GC36542A.

49. Kang, L., Xu, P., Zhang, B., Tsai, H., Han, X., and Wang, H.-L. Laser wavelength- and power-dependent plasmon-driven chemical reactions monitored using single particle surface enhanced Raman spectroscopy. Chem. Commun. 2013, 49(33), 3389–3391. https://doi.org/10.1039/ C3CC40732B.

50. Kaur, G., Tanwar, S., Kaur, V., Biswas, R., Saini, S., Haldar, K. K., and Sen, T. Interfacial design of gold/silver core–shell nanostars for plasmonenhanced photocatalytic coupling of 4-aminothiophenol. J. Mater. Chem. C 2021, 9(42), 15284–15294. https://doi.org/10.1039/D1TC03733A. 51. Koopman, W., Sarhan, R. M., Stete, F., Schmitt, C. N. Z., and Bargheer, M. Decoding the kinetic limitations of plasmon catalysis: The case of 4-nitrothiophenol dimerization. Nanoscale 2020, 12(48), 24411– 24418. https://doi.org/10.1039/D0NR06039A.

52. Cho, F.-H., Kuo, S.-C., and Lai, Y.-H. Surface-plasmon-induced azo coupling reaction between nitro compounds on dendritic silver monitored by surface-enhanced raman spectroscopy. RSC Adv. 2017, 7(17), 10259–10265. https://doi.org/10.1039/C7RA00374A. 53. Hartman, T., Wondergem, C. S., Kumar, N., van den Berg, A., and Weckhuysen, B. M. Surface- and tip-enhanced raman spectroscopy in catalysis. J. Phys. Chem. Lett. 2016, 7(8), 1570–1584. https://doi. org/10.1021/acs.jpclett.6b00147.

54. Zhang, Z., Kinzel, D., and Deckert, V. Photo-induced or plasmon-induced reaction: Investigation of the light-induced azo-coupling of amino groups. J. Phys. Chem. C 2016, 120(37), 20978–20983. https://doi. org/10.1021/acs.jpcc.6b03233.

References

55. Huang, Y.-F., Wu, D.-Y., Zhu, H.-P., Zhao, L.-B., Liu, G.-K., Ren, B., and Tian, Z.-Q. Surface-enhanced raman spectroscopic study of p-aminothiophenol. Phys. Chem. Chem. Phys. 2012, 14(24), 8485–8497. https://doi.org/10.1039/C2CP40558J. 56. Ren, X., Tan, E., Lang, X., You, T., Jiang, L., Zhang, H., Yin, P., and Guo, L. Observing reduction of 4-nitrobenzenthiol on gold nanoparticles in situ using surface-enhanced Raman spectroscopy. Phys. Chem. Chem. Phys. 2013, 15(34), 14196–14201. https://doi.org/10.1039/C3CP51385H.

57. Pradhan, N., Pal, A., and Pal, T. Silver nanoparticle catalyzed reduction of aromatic nitro compounds. Colloids Surf. Physicochem. Eng. Asp. 2002, 196(2), 247–257. https://doi.org/10.1016/S0927-7757(01)01040-8. 58. Wu, W., Lei, M., Yang, S., Zhou, L., Liu, L., Xiao, X., Jiang, C., and Roy, V. A. L. A one-pot route to the synthesis of alloyed Cu/Ag bimetallic nanoparticles with different mass ratios for catalytic reduction of 4-nitrophenol. J. Mater. Chem. A 2015, 3(7), 3450–3455. https://doi. org/10.1039/C4TA06567K.

59. Han, Q., Zhang, C., Gao, W., Han, Z., Liu, T., Li, C., Wang, Z., He, E., and Zheng, H. Ag-Au alloy nanoparticles: Synthesis and in situ monitoring SERS of plasmonic catalysis. Sens. Actuators B-Chem. 2016, 231, 609– 614. https://doi.org/10.1016/j.snb.2016.03.068.

60. Shah, D. and Kaur, H. Resin-trapped gold nanoparticles: An efficient catalyst for reduction of nitro compounds and Suzuki–Miyaura coupling. J. Mol. Catal. Chem. 2014, 381, 70–76. https://doi.org/10.1016/j. molcata.2013.10.004.

61. Liu, H., Yu, J., Liu, Y., and Liu, Y. Plasmonic acceleration of nitrophenol reduction upon catalysis by robust gold nanoparticle donut assemblies. Catal. Commun. 2020, 146, 106137. https://doi.org/10.1016/j. catcom.2020.106137. 62. Zeng, J., Zhang, Q., Chen, J., and Xia, Y. A comparison study of the catalytic properties of Au-based nanocages, nanoboxes, and nanoparticles. Nano Lett. 2010, 10(1), 30–35. https://doi.org/10.1021/nl903062e.

63. Wu, X.-Q., Wu, X.-W., Huang, Q., Shen, J.-S., and Zhang, H.-W. In situ synthesized gold nanoparticles in hydrogels for catalytic reduction of nitroaromatic compounds. Appl. Surf. Sci. 2015, 331, 210–218. https:// doi.org/10.1016/j.apsusc.2015.01.077. 64. Cui, Q., Xia, B., Mitzscherling, S., Masic, A., Li, L., Bargheer, M., and Möhwald, H. Preparation of gold nanostars and their study in selective catalytic reactions. Colloids Surf. Physicochem. Eng. Asp. 2015, 465, 20– 25. https://doi.org/10.1016/j.colsurfa.2014.10.028.

89

90

Surface Plasmon Resonance‒Mediated Organic Synthesis

65. Haldar, K. K., Kundu, S., and Patra, A. Core-size-dependent catalytic properties of bimetallic Au/Ag core–shell nanoparticles. ACS Appl. Mater. Interfaces 2014, 6(24), 21946–21953. https://doi.org/10.1021/ am507391d.

66. Xiao, Q., Sarina, S., Waclawik, E. R., Jia, J., Chang, J., Riches, J. D., Wu, H., Zheng, Z., and Zhu, H. Alloying gold with copper makes for a highly selective visible-light photocatalyst for the reduction of nitroaromatics to anilines. ACS Catal. 2016, 6(3), 1744–1753. https://doi.org/10.1021/ acscatal.5b02643.

67. Hajfathalian, M., Gilroy, K. D., Yaghoubzade, A., Sundar, A., Tan, T., Hughes, R. A., and Neretina, S. Photocatalytic enhancements to the reduction of 4-nitrophenol by resonantly excited triangular gold– copper nanostructures. J. Phys. Chem. C 2015, 119(30), 17308–17315. https://doi.org/10.1021/acs.jpcc.5b04618. 68. He, R., Wang, Y.-C., Wang, X., Wang, Z., Liu, G., Zhou, W., Wen, L., Li, Q., Wang, X., Chen, X., Zeng, J., and Hou, J. G. Facile synthesis of pentacle gold–copper alloy nanocrystals and their plasmonic and catalytic properties. Nat. Commun. 2014, 5(1), 4327. https://doi.org/10.1038/ ncomms5327. 69. Vysakh, A. B., Babu, C. L., and Vinod, C. P. Demonstration of synergistic catalysis in Au@Ni bimetallic core–shell nanostructures. J. Phys. Chem. C 2015, 119(15), 8138–8146. https://doi.org/10.1021/jp5128089.

70. Li, J., Long, Y., Liu, Y., Zhang, L., Wang, Q., Wang, X., Song, S., and Zhang, H. Robust synthesis of gold-based multishell structures as plasmonic catalysts for selective hydrogenation of 4-nitrostyrene. Angew. Chem. Int. Ed. 2020, 59(3), 1103–1107. https://doi.org/10.1002/ anie.201910836. 71. Verma, A., Shukla, M., Kumar, S., Pal, S., and Sinha, I. Mechanism of visible light enhanced catalysis over curcumin functionalized Ag nanocatalysts. Spectrochim. Acta. A. Mol. Biomol. Spectrosc. 2020, 240, 118534. https://doi.org/10.1016/j.saa.2020.118534.

72. Barbosa, E. C. M., Fiorio, J. L., Mou, T., Wang, B., Rossi, L. M., and Camargo, P. H. C. Reaction pathway dependence in plasmonic catalysis: Hydrogenation as a model molecular transformation. Chem. – Eur. J. 2018, 24(47), 12330–12339. https://doi.org/10.1002/ chem.201705749.

73. Yin, H., Kuwahara, Y., Mori, K., and Yamashita, H. RuPd alloy nanoparticles supported on plasmonic HxMoO3–y for efficient photocatalytic reduction of p-nitrophenol. Eur. J. Inorg. Chem. 2019, 2019(33), 3745–3752. https://doi.org/10.1002/ejic.201900801.

References

74. Tong, F., Liang, X., Ma, F., Bao, X., Wang, Z., Liu, Y., Wang, P., Cheng, H., Dai, Y., Huang, B., and Zheng, Z. Plasmon-mediated nitrobenzene hydrogenation with formate as the hydrogen donor studied at a single-particle level. ACS Catal. 2021, 11(7), 3801–3809. https://doi. org/10.1021/acscatal.1c00164.

75. Hao, C.-H., Guo, X.-N., Pan, Y.-T., Chen, S., Jiao, Z.-F., Yang, H., and Guo, X.-Y. Visible-light-driven selective photocatalytic hydrogenation of cinnamaldehyde over Au/SiC catalysts. J. Am. Chem. Soc. 2016, 138(30), 9361–9364. https://doi.org/10.1021/jacs.6b04175. 76. Alejo, C. J. B., Fasciani, C., Grenier, M., Netto-Ferreira, J. C., and Scaiano, J. C. Reduction of resazurin to resorufin catalyzed by gold nanoparticles: Dramatic reaction acceleration by laser or LED plasmon excitation. Catal. Sci. Technol. 2011, 1(8), 1506–1511. https://doi.org/10.1039/ c1cy00236h. 77. Xu, W., Kong, J. S., Yeh, Y.-T. E., and Chen, P. Single-molecule nanocatalysis reveals heterogeneous reaction pathways and catalytic dynamics. Nat. Mater. 2008, 7(12), 992–996. https://doi.org/10.1038/nmat2319.

78. Zhou, X., Xu, W., Liu, G., Panda, D., and Chen, P. Size-dependent catalytic activity and dynamics of gold nanoparticles at the single-molecule level. J. Am. Chem. Soc. 2010, 132(1), 138–146. https://doi.org/10.1021/ ja904307n. 79. Chen, P., Zhou, X., Andoy, N. M., Han, K.-S., Choudhary, E., Zou, N., Chen, G., and Shen, H. Spatiotemporal catalytic dynamics within single nanocatalysts revealed by single-molecule microscopy. Chem. Soc. Rev. 2014, 43(4), 1107–1117. https://doi.org/10.1039/C3CS60215J. 80. Quiroz, J., Barbosa, E. C. M., Araujo, T. P., Fiorio, J. L., Wang, Y.-C., Zou, Y.-C., Mou, T., Alves, T. V., de Oliveira, D. C., Wang, B., Haigh, S. J., Rossi, L. M., and Camargo, P. H. C. Controlling reaction selectivity over hybrid plasmonic nanocatalysts. Nano Lett. 2018, 18(11), 7289–7297. https:// doi.org/10.1021/acs.nanolett.8b03499.

81. Wang, R., Liu, H., Wang, X., Li, X., Gu, X., and Zheng, Z. Plasmonenhanced furfural hydrogenation catalyzed by stable carbon-coated copper nanoparticles driven from metal–organic frameworks. Catal. Sci. Technol. 2020, 10(19), 6483–6494. https://doi.org/10.1039/ D0CY01162B.

82. Garcia-Garcia, I., Lovell, E. C., Wong, R. J., Barrio, V. L., Scott, J., Cambra, J. F., and Amal, R. Silver-based plasmonic catalysts for carbon dioxide reduction. ACS Sustain. Chem. Eng. 2020, 8(4), 1879–1887. https://doi. org/10.1021/acssuschemeng.9b06146.

91

92

Surface Plasmon Resonance‒Mediated Organic Synthesis

83. Fasciani, C., Alejo, C. J. B., Grenier, M., Netto-Ferreira, J. C., and Scaiano, J. C. High-temperature organic reactions at room temperature using plasmon excitation: Decomposition of dicumyl peroxide. Org. Lett. 2011, 13(2), 204–207. https://doi.org/10.1021/ol1026427.

84. Bakhtiari, A. B. S., Hsiao, D., Jin, G., Gates, B. D., and Branda, N. R. An efficient method based on the photothermal effect for the release of molecules from metal nanoparticle surfaces. Angew. Chem.-Int. Ed. 2009, 48(23), 4166–4169. https://doi.org/10.1002/anie.200805303. 85. Liang, C., Zhang, Y., Zhang, B., Liu, X.-M., Gao, G.-L., Cao, J., and Xu, P. Plasmonic heating-promoted photothermal synthesis of α-cyanoacrylonitriles over Au/h-BN catalysts. Front. Chem. 2021, 9. 86. Han, D., Bao, Z., Xing, H., Yang, Y., Ren, Q., and Zhang, Z. Fabrication of plasmonic Au–Pd alloy nanoparticles for photocatalytic Suzuki– Miyaura reactions under ambient conditions. Nanoscale 2017, 9(18), 6026–6032. https://doi.org/10.1039/C7NR01950E.

87. Wang, F., Li, C., Chen, H., Jiang, R., Sun, L.-D., Li, Q., Wang, J., Yu, J. C., and Yan, C.-H. Plasmonic harvesting of light energy for Suzuki coupling reactions. J. Am. Chem. Soc. 2013, 135(15), 5588–5601. https://doi. org/10.1021/ja310501y. 88. Yoshii, T., Kuwahara, Y., Mori, K., and Yamashita, H. Design of Pd– graphene–Au nanorod nanocomposite catalyst for boosting Suzuki– Miyaura coupling reaction by assistance of surface plasmon resonance. J. Phys. Chem. C 2019, 123(40), 24575–24583. https://doi.org/10.1021/ acs.jpcc.9b06609.

89. Zhu, X., Jia, H., Zhu, X.-M., Cheng, S., Zhuo, X., Qin, F., Yang, Z., and Wang, J. Selective Pd deposition on Au nanobipyramids and Pd site-dependent plasmonic photocatalytic activity. Adv. Funct. Mater. 2017, 27(22), 1700016. https://doi.org/10.1002/adfm.201700016.

90. Sarhid, I., Abdellah, I., Martini, C., Huc, V., Dragoe, D., Beaunier, P., Lampre, I., and Remita, H. Plasmonic catalysis for the Suzuki–Miyaura cross-coupling reaction using palladium nanoflowers. New J. Chem. 2019, 43(11), 4349–4355. https://doi.org/10.1039/C8NJ06370B.

91. Koohgard, M. and Hosseini-Sarvari, M. Enhancement of Suzuki–Miyaura coupling reaction by photocatalytic palladium nanoparticles anchored to TiO2 under visible light irradiation. Catal. Commun. 2018, 111, 10– 15. https://doi.org/10.1016/j.catcom.2018.03.026. 92. Gangishetty, M. K., Fontes, A. M., Malta, M., Kelly, T. L., and Scott, R. W. J. Improving the rates of Pd-catalyzed reactions by exciting the surface plasmons of AuPd bimetallic nanotriangles. RSC Adv. 2017, 7(64), 40218–40226. https://doi.org/10.1039/C7RA07264C.

References

93. Nemygina, N. A., Nikoshvili, L. Zh., Tiamina, I. Yu., Bykov, A. V., Smirnov, I. S., LaGrange, T., Kaszkur, Z., Matveeva, V. G., Sulman, E. M., and KiwiMinsker, L. Au core–Pd shell bimetallic nanoparticles immobilized within hyper-cross-linked polystyrene for mechanistic study of Suzuki cross-coupling: Homogeneous or heterogeneous catalysis? Org. Process Res. Dev. 2018, 22(12), 1606–1613. https://doi.org/10.1021/acs. oprd.8b00272. 94. Rohani, S., Ziarati, A., Ziarani, G. M., Badiei, A., and Burgi, T. Engineering of highly active Au/Pd supported on hydrogenated urchin-like yolk@shell TiO2 for visible light photocatalytic Suzuki coupling. Catal. Sci. Technol. 2019, 9(14), 3820–3827. https://doi.org/10.1039/C9CY00798A. 95. Li, Z. and Kurouski, D. Probing the plasmon-driven Suzuki–Miyaura coupling reactions with cargo-TERS towards tailored catalysis. Nanoscale 2021, 13(27), 11793–11799. https://doi.org/10.1039/ D1NR02478G.

96. Li, Z., Gong, Y., Zhang, X., Wen, Y., Yao, J., Hu, M., He, M., Liu, J., Li, R., Wang, F., and Zhang, C. Plasmonic coupling-enhanced in situ photothermal nanoreactor with shape selective catalysis for C–C coupling reaction. Nano Res. 2020, 13(10), 2812–2818. https://doi.org/10.1007/s12274020-2933-3. 97. Huynh, L. T. M., Trinh, H. D., Lee, S., and Yoon, S. Plasmon-driven protodeboronation reactions in nanogaps. Nanoscale 2020. https:// doi.org/10.1039/D0NR07023H.

98. Kumar, D., Lee, A., Lee, T., Lim, M., and Lim, D.-K. Ultrafast and efficient transport of hot plasmonic electrons by graphene for Pt free, highly efficient visible-light responsive photocatalyst. Nano Lett. 2016, 16(3), 1760–1767. https://doi.org/10.1021/acs.nanolett.5b04764. 99. Li, Y., Wen, M., Wang, Y., Tian, G., Wang, C., and Zhao, J. Plasmonic hot electrons from oxygen vacancies for infrared light-driven catalytic CO2 reduction on Bi2O3−x. Angew. Chem. Int. Ed. https://doi.org/10.1002/ anie.202010156. 100. Zhang, H., Itoi, T., Konishi, T., and Izumi, Y. Dual photocatalytic roles of light: Charge separation at the band gap and heat via localized surface plasmon resonance to convert CO2 into CO over silver-zirconium oxide. J. Am. Chem. Soc. 2019, 141(15), 6292–6301. https://doi.org/10.1021/ jacs.8b13894. 101. González-Béjar, M., Peters, K., Hallett-Tapley, G. L., Grenier, M., and Scaiano, J. C. Rapid one-pot propargylamine synthesis by plasmon

93

94

Surface Plasmon Resonance‒Mediated Organic Synthesis

mediated catalysis with gold nanoparticles on ZnO under ambient conditions. Chem. Commun. 2013, 49(17), 1732–1734. https://doi. org/10.1039/C3CC38287G.

102. Zhao, J., Zheng, Z., Bottle, S., Chou, A., Sarina, S., and Zhu, H. Highly efficient and selective photocatalytic hydroamination of alkynes by supported gold nanoparticles using visible light at ambient temperature. Chem. Commun. 2013, 49(26), 2676–2678. https://doi. org/10.1039/C3CC38985E.

103. Meng, J., Xiao, Y., Huang, Y., Liu, X., and Zhao, J. Solvent-free hydroamination of phenylacetylene by plasmonic gold nanoparticles coupled with a TiO2 2D photonic layer on nanotube arrays. New J. Chem. 2020, 44(40), 17289–17295. https://doi.org/10.1039/D0NJ04132G. 104. Pineda, A., Gomez, L., Balu, A. M., Sebastian, V., Ojeda, M., Arruebo, M., Romero, A. A., Santamaria, J., and Luque, R. Laser-driven heterogeneous catalysis: Efficient amide formation catalysed by Au/SiO2 systems. Green Chem. 2013, 15(8), 2043–2049. https://doi.org/10.1039/ C3GC40166A.

105. Bazyar, Z. and Hosseini-Sarvari, M. Au@ZnO core–shell: Scalable photocatalytic trifluoromethylation using CF3CO2Na as an inexpensive reagent under visible light irradiation. Org. Process Res. Dev. 2019, 23(11), 2345–2353. https://doi.org/10.1021/acs.oprd.9b00225.

106. Sun, X., Zou, Y., and Jiang, J. Surface plasmon resonances enhanced click chemistry through synergistic photothermal and hot electron effects. Chem. Commun. 2019, 55(33), 4813–4816. https://doi.org/10.1039/ C9CC01456J.

107. Guselnikova, O., Váňa, J., Phuong, L. T., Panov, I., Rulíšek, L., Trelin, A., Postnikov, P., Švorčík, V., Andris, E., and Lyutakov, O. Plasmon-assisted click chemistry at low temperature: An inverse temperature effect on the reaction rate. Chem. Sci. 2021, 12(15), 5591–5598. https://doi. org/10.1039/D0SC05898J. 108. Tijunelyte, I., Guenin, E., Lidgi-Guigui, N., Colas, F., Ibrahim, J., Toury, T., and de la Chapelle, M. L. Nanoplasmonics tuned “click chemistry.” Nanoscale 2016, 8(13), 7105–7112. https://doi.org/10.1039/ C5NR09018K. 109. Dos Santos, C. G., Marquez, D. T., Crites, C.-O. L., Netto-Ferreira, J. C., and Scaiano, J. C. Plasmon heating mediated Friedel-Crafts alkylation of anisole using supported AuNP@Nb2O5 catalysts. Tetrahedron Lett. 2017, 58(5), 427–431. https://doi.org/10.1016/j.tetlet.2016.12.048.

References

110. Teixeira, I. F., Homsi, M. S., Geonmonond, R. S., Rocha, G. F. S. R., Peng, Y.-K., Silva, I. F., Quiroz, J., and Camargo, P. H. C. Hot electrons, hot holes, or both? Tandem synthesis of imines driven by the plasmonic excitation in Au/CeO2 nanorods. Nanomaterials 2020, 10(8), 1530. https://doi. org/10.3390/nano10081530. 111. Wang, X., Wang, R., Wang, J., Fan, C., and Zheng, Z. The synergistic role of the support surface and Au–Cu alloys in a plasmonic Au–Cu@LDH photocatalyst for the oxidative esterification of benzyl alcohol with methanol. Phys. Chem. Chem. Phys. 2020, 22(3), 1655–1664. https:// doi.org/10.1039/C9CP05992.

95

Chapter 4

Plasmon-Mediated Homogeneous Catalysis

Léa Gimeno,a Somayeh Talebzadeh,b Scott Trammell,c Clémence Queffélec,a and D. Andrew Knightc aUniversity

of Nantes, CNRS, CEISAM, UMR 6230, F-44000 Nantes, France Institute of Technology, 150 West University Blvd, Melbourne, FL 32901, USA cUS Naval Research Laboratory, 4555 Overlook Avenue, Washington DC 20375, USA [email protected] bFlorida

4.1 Introduction Homogeneous catalytic reactions, especially those involving transition metals and organometallic reagents, have been described in the academic literature and used in industrial-scale reactions for over 60 years. Classic examples include the Wacker process for the oxidation of ethylene to acetaldehyde [1], the rhodiumcatalyzed hydrogenation of alkenes as described by Wilkinson [2], the Sharpless epoxidation reaction [3], and more recently palladium-catalyzed cross-coupling reactions for which Suzuki, Heck, and Negishi received the Nobel Prize for Chemistry in 2010 [4]. Traditional homogeneous catalysis has now been extended Plasmonic Nanomaterials: Characterization and Applications in Organic Synthesis and Catalysis Edited by Clémence Queffélec and D. Andrew Knight Copyright © 2024 Jenny Stanford Publishing Pte. Ltd. ISBN 978-981-5129-09-0 (Hardcover), 978-1-003-47406-7 (eBook) www.jennystanford.com

98

Plasmon-Mediated Homogeneous Catalysis

to include reactions that take place in alternative solvents such as water and ionic liquids [5], enantioselective catalysis [6], and supported molecular catalysis [7], which reflects combining the advantages of homogeneous and heterogenous catalysis. Some of the well-documented advantages of homogeneous catalysis over its heterogeneous counterpart include ease of separation of product from reaction mixture, mild operating conditions, improved selectivities, and amenability toward mechanistic understanding [8]. Although the application of surface plasmon resonance (SPR) to catalytic reactions has been dominated by heterogeneous catalytic systems [9], the exploitation of the localized surface plasmon resonance (LSPR) effect exhibited by certain metal nanoparticles to homogeneous catalysis has started to emerge. Using the design principle in which a molecular homogeneous catalyst, e.g., a ligandstabilized transition metal complex, is anchored to the surface of a plasmonically active metal nanoparticle, energy can be transferred from the light source, e.g., laser or LED, to the plasmonic support. This energy associated with the light absorption can be utilized for (a) energy transfer sensitization, (b) redox processes of nearby molecule, e.g., catalyst or substrate, and (c) localized heating of the microenvironment, which results in the acceleration of reaction rate (Fig. 4.1).

Figure 4.1 Pathways for metal nanoparticle energy coupling to local environment (from Ref. [10]).

Although photochemical homogeneous catalytic reactions have been known for some time [11], the use of LSPR represents a new direction in photo-assisted catalysis. This chapter describes the use of LSPR for the acceleration of homogeneous catalytic reactions based on a recent literature survey with an emphasis on degradation and synthesis.

Catalytic Hydrolysis

4.2 Catalytic Hydrolysis 4.2.1 Hydrolysis of Phosphate Esters The catalytic degradation or cleavage of phosphate and thiophosphate esters, especially those involving a metal-mediated hydrolytic mechanism, is of interest as an approach for the development of materials suitable for the destruction of chemical warfare agents and toxic pesticides as well as selective catalysts for the breakdown of DNA and RNA (artificial endonucleases) [12]. Catalysts based on copper(II) and lanthanides have been some of the most successful. The role of metal ions and metal–ion complexes stabilized with organic ligands such as 2,2¢-bipyridine (bpy) and 1,10-phenanthroline in the hydrolytic catalysis of phosphate esters has been known for some time, and the mechanism of these reactions has been extensively studied [13]. The accepted mechanism is that the phosphate ester initially displaces a coordinated water molecule and binds to the metal center. Intramolecular attack by a second coordinated water ligand occurs. Scheme 4.1 shows an example with the Cu(bpy)2+(aq) complex. These reactions suffer from slow reaction kinetics and high catalyst/substrate ratios and as such are highly amenable to kinetic enhancement using LSPR. In 2012, one of the first homogeneous metal-based catalytic reactions accelerated by SPR was reported by Trammell et al. [15]. The authors described a 10 nm gold nanoparticle (AuNP) platform, which supported a copper(II) bipyridine complex via stable Au-thiol interactions. The supported complex was used for the hydrolysis of methyl parathion, a broad-spectrum thiophosphate insecticide, which is environmentally persistent. Using a lowpower (120 mW), 532 nm green laser, it was demonstrated that a rate acceleration of ¥2-fold occurred for the hydrolysis of methyl parathion by irradiating the plasmon absorption band of AuNPs, which were capped with the copper-bipyridine catalyst. The phenomenon of the acceleration of the initial rate of hydrolysis of MeP by the excitation of the plasmon absorption band of the AuNPs could have several mechanistic explanations. For example, energy transfer sensitizing the Cu metal center could create an excited d–d state, which might accelerate the rate of hydrolysis if a dissociative step at the metal center is rate limiting. Another

99

Cu

H

O

O

H

H

N

N

+

Cu

-O

O

O

H

OR

P

O

-O

P

OR

R

OR

O-

N

N

N

N

Cu

Cu

OR

O

OH

H

O

O

P

H

H

OR

O

+

+

H

-O

N

N

O

P

O

Cu

Scheme 4.1 The general mechanism of hydrolytic catalysis of phosphate esters by Cu(bpy)2+(aq) (from Ref. [14]).

N

N

+

OR

O

O

H

P OR

OR

O-

100 Plasmon-Mediated Homogeneous Catalysis

Catalytic Carbon–Carbon Bond Formation

explanation maybe the reduction of Cu(II) to Cu(I) by the AuNPs, creating a more active catalyst. However, the simplest explanation is the creation of localized heating of microenvironment near the AuNP surface. Since we see an increase in rate by a factor of ¥2, an increase of 10 degrees in the microenvironment would be a common prediction by the Arrhenius equation for this type of reaction. AuNPs are good light-to-heat converters with excitation of their plasmon absorption bands dissipating energy into the microenvironment.

4.3 Catalytic Carbon–Carbon Bond Formation 4.3.1 Ullmann Reaction

Forato et al. [16] developed copper-polypyridine catalysts grafted onto core–shell silver nanomaterial and used this plasmonic active nanocatalyst in catalysis. They focused on the formation of C–C bonds, and the model reaction studied was the Ullmann reaction, which is a classic example of the catalytic carbon–carbon bond-forming process. This reaction involves the coupling of aryl halides mediated by copper (Scheme 4.2). 2

Cu source

X

R

base

R

R

X = I or Br and R = NO2, OMe .... Scheme 4.2 Example of the Ullmann reaction catalyzed by copper.

In that regard, the ligand featuring a 2,2¢-bipyridine backbone bearing one terminal phosphonic acid tether was obtained in five steps with a 30% overall yield (Scheme 4.3). N

N

SeO2, dioxane

N

N

24 h, reflux + H2N

O

1) EtOH, rt

2) TMSBr, then MeOH CHO quantitative O P OEt OEt

Scheme 4.3 Synthesis of the targeted ligand L.

N

N

N O

Ligand L

PO3H2

101

102

Plasmon-Mediated Homogeneous Catalysis

Core–shell nanoparticles were synthesized using adapted protocols described in the literature [17, 18]. Reduction of silver nitrate with hydrazine monohydrate led to the formation of AgNPs, and the TiO2 shell was then formed through the hydrolysis of titanium isopropoxide in water, leading to Ag@TiO2 NPs. From Fig. 4.2a, it can be seen that AgNPs are randomly embedded in a TiO2 matrix. Then, a water solution of ligand L was added to the Ag@TiO2 NPs, which led to the grafting of the ligand onto the NPs to form Ag@TiO2@L NPs (Fig. 4.2b). Finally, an aqueous solution of Cu(CH3CN)4PF6 was added to the Ag@TiO2@L NPs and the final nanocatalyst Ag@ TiO2@L@Cu(I) was obtained (Fig. 4.2c) as spherical nanoparticles with an average diameter of 19 nm.

a

b

c

Figure 4.2 TEM images of (a) Ag@TiO2 NPs, (b) Ag@TiO2@L NPs, (c) Ag@ TiO2@L-Cu(I) nanoparticles.

UV–vis spectra were recorded to confirm the presence of the plasmonic band that was centered around 445 nm for the nanocatalyst. Besides, the SPR band is quite large, which allows the use of various wavelength for the catalytic reaction irradiation, especially 450 nm or 488 nm wavelengths. This nanocatalyst was further characterized by surfaceenhanced Raman spectroscopy (SERS) [19] and X-ray absorption (XANES and EXAFS) [20]. Thus, the oxidation state of the copper ion was confirmed by XANES and the copper complex geometry grafted on the surface by EXAFS and SERS. The best EXAFS model suggested a low coordinate copper(I) geometry that included one bpy ligand with two nitrogen atoms at 1.90(1) Å and an elongated distance to a second light atom (i.e., C, N, or O) at 2.04(2) Å. Besides, by SERS, different peaks with a small shift in the wavenumber values were attributed to different modes of vibration of bipyridine and confirming that copper was coordinated to the bipyridine skeleton.

Table 4.1

Ullmann coupling reaction under different conditions

Catalyst

Laser Wavelength

Power

Time

Temperature

Yield

1

CuI (5%) + 2,2¢-bipyridine (15%)

No laser



24 h

110°C

70%

3

Ag@TiO2

488 nm

220 mW

8h

T.A.

No conversion

Ag@TiO2@L@-Cu(I) (≈ 1%)

488 nm

220 mW

8h

T.A.

65%

2

4

5

Ag@TiO2@L@-Cu(I) (≈ 1%)

Ag@TiO2@- L@Cu(I) (≈ 1%)

No laser 488 nm



220 mW

24 h 5h

110°C T.A.

= |g> | g >, |f> = |g> |t >. | g > and |t > are different vibrational states of the molecule; |g> is the ground electronic states of the nanoparticle. In step 1, the nanoparticle is excited from |g> to a state |h> and then de-excited back to |g>, transferring energy to the molecule. In step 3, the nanoparticle is excited from |g> to a state |h¢> by receiving energy from the molecule and then de-excited back to |g> by emitting the scattering photon. The plasmon excited states |h> and |h¢> differ by the vibrational quantum of energy deposited in the molecule (going from | g > to | t >), which is also the energy difference between the incident and scattered photons. Adapted from Ref. [23].

A different optical process arises if the metal surface is covered by adsorbate molecules: The energy stored in the metal as LSP

115

116

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

may be transferred via dipole–dipole interaction to the molecule. Vibrational modes may be activated before a further energy transfer occurs, this time from the molecule back to the metal. Now the scattered photon has a different energy, and the difference in energy is the quantity gained or lost by the molecule through its vibration. This is an inelastic scattering or Raman process (involving both the molecule and the metal surface). As reported in the diagram, the whole process may be figuratively divided into three resonant steps: Most importantly, the LSP resonances take part twice (but with slightly different energies, hninc and hnsc); they are twice excited and twice de-excited. Compared with the normal resonance Raman process, where the molecule interacts directly with the incident electromagnetic field, the field experienced by the adsorbed molecule is the local field enhanced by the induced oscillations of the electrons in the metal particle. All optical processes may profit from this local field enhancement, and surface-enhanced infrared spectroscopy (SEIRA) and metalenhanced fluorescence (MEF) exist as well. These are single-photon absorption processes, and the signal enhancement is “only” the square of the field amplification, since there is only one photon–LSP interaction. The substantially greater popularity of SERS lies in the fourth power relationship stemming from the LSP excitation taking part twice in the overall optical scattering process: First, the incident field (photon) interacts with the LSP with an energy hninc, and then the scattered field at the Raman-shifted frequency hnsc interacts with the LSP (Figs. 5.2 and 5.3). Indeed, the small difference in energy between hninc and hnsc (a molecular vibrational quantum compared to the electronic energies of the LSP resonances) is the reason why the SERS fourth power law is an approximation. More precisely one should multiply the amplification of the field at the energy hninc (the square of this field intensity) by the amplification of the field at the energy hnsc (the square of this different field intensity). This is approximated by considering the same amplification at the two close energies and for the Raman signal enhancement |Einc|2*|Esc|2 ~ |Einc|4. Hot spots are particular locations on the surface where the electromagnetic field is strongly enhanced due to the geometry of the corresponding feature (remember that the intensity of the LSP resonances depends on the size, shape, and metal dielectric

Surface-Enhanced Raman Spectroscopy

constant; see Chapter 1 for more details). A simple example is the case of elongated nanoparticles (nanorods) illuminated with light polarized along the long axis. In this case, the strongest electron oscillations and the strongest field intensity are localized near the apexes, and the molecules placed near the apexes will provide a much stronger SERS signal than the molecules placed near the sides of the rod. The concept of hot spots is reminiscent of catalytic active sites: On a catalytic surface, a few sites are mostly responsible for the catalytic activity, the rest of the surface being rather inactive. Similarly, the molecules adsorbed at the hot spots are responsible for most of the SERS signal, while all the other molecules, despite being illuminated by the laser, provide only a normal, unenhanced Raman signal, almost negligible with respect to the SERS. Thus, the SERS signal originates from a much smaller number of molecules than the number that can be estimated by taking in account the laser spot size and the illuminated geometrical area. Furthermore, the enhanced electromagnetic field expands only up to ~100 nm away from the surface, meaning that the volume sensed by SERS is very limited. There is a consensus on the magnitude and spatial decay of the enhanced electromagnetic field intensity with the distance from the nanostructure: The strongest enhanced fields decay the fastest (within 10 nm); the weaker fields, generated by larger structures, decay more slowly. This is very relevant for SERS, since an exceedingly small number of molecules located next to a very (and small) hot spot contribute to the vast majority of the SERS signal; the vast majority of molecules located in “colder” regions do not contribute much. This is the tenet for the very high SERS sensitivity and its potential for single-molecule detection. Concerning the shape and size of these hot spots, other than sharp protrusions (with end radius 104. Adapted from Ref. [37].

Tip-Enhanced Raman Spectroscopy



C=

I tip down I tip up

=

Inearfield + Ifar - field Ifar - field

(5.2)

Ê Inearfield ˆ 2 ÁË Rspot Anearfield ˜¯ Afar - field = (C - 1) EF = = (C - 1) 2 (5.3) Anearfield Ê Ifar - field ˆ Rtip ÁË ˜ Afar - field ¯

A is the focal area defined by the NA of the objective and the employed wavelength, and R is its radius. To obtain the enhancement factor, the tip-down and tip-up intensities should be normalized by the number of molecules participating in the respective spectra; for a uniformly distributed molecular layer of Raman scatterer, the farfield and nearfield areas can be considered. The far-field is the laser spot area (having usually circular spots of radius Rspot ~0.5 mm and bottom illumination); all the illuminated molecules participate in the “unenhanced” far-field signal. The nearfield area is the small disk region below the tip, having a radius approximated by the physical radius of the apex (Rtip ~30 nm). Only a very small subset of all the illuminated molecules contributes to the TERS signal, but this smaller number of molecules provides a much stronger Raman signal than all the illuminated molecules thanks to the tip enhancement. In the gap-mode TERS, where the apex and metal substrate form a nanocavity, the effective lateral extension (localization) of the squeezed plasmon mode is L = (2RtipD)1/2 [38], where D is the distance from the flat surface to the bottom of a sphere having a radius Rtip and the TERS radius can be approximated to L/2. Note that the tip-down spectra are an incoherent sum of the far-field and nearfield signals; in a number of cases, the far-field is extremely weak compared to the TERS intensity or altogether absent and the tip-up spectra consisting only of a baseline due to the substrate; in this case, formula 5.3 can be simplified, removing the factor one. Similar formulas can be used for SERS if the number and size of the contributing hot spots are known, as it is indeed well known for TERS or for certain lithographically fabricated SERS substrates. Thus, the estimations on the absolute Raman signal enhancement factor in SERS are more varied and based on assumptions, than in the case of TERS. Generally, apart from the very different intensity level, a TERS spectrum mostly resembles the unenhanced, far-field spectrum of the adsorbed molecules. In some

129

130

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

instances, band narrowing and maximum shifts are to be ascribed to TERS, probing a much smaller number of molecules with less averaging of different surface sites and molecular configurations. While Fig. 5.6 shows the signal enhancement, the first advantageous TERS feature, Fig. 5.7 nicely shows the improvement in lateral resolution when passing from far-field Raman imaging to nearfield Raman imaging [39]. The same isolated single-wall carbon nanotube (SWCNT) is probed: The left image corresponds to a classical confocal Raman image (no illuminated tip), specifically to the intensity of the 2D band. SWCNTs provide strong Raman scattering due to resonant effects; thus, even a single isolated tube yields sufficient signal without any other external enhancement. Anyhow, the limit of far-field imaging is evident considering that SWCNTs are almost 1D cylindrical objects with typical lateral size of ~1 nm. In the confocal Raman image, the tube is seen with a much larger width of ~350 nm. This apparent width is due to the convolution of the real width (almost negligible) of SWCNTs and the diffraction-limited spot size. In the TERS image (same optically scanned region, but with the tip brought in contact with the sample and illuminated), the apparent width is only ~14 nm, since now the real width is convoluted with the size of the nearfield (TERS) radius, which also defines the optical resolution in this particular experiment.

Figure 5.7 Raman confocal image (left) and TERS image of the same SWCNT deposited on a glass slide. The color reflects the intensity of the SWCNT 2D band at 2640 cm−1: A higher intensity is represented by a brighter color. Reproduced with permission from Ref. [39].

Tip-Enhanced Raman Spectroscopy

Figure 5.8 TERS spectra recorded from a single H2TBPP molecule on Ag (111). The blue and red traces refer to two different positions of the tip on the molecule; the black trace is relative to the tip positioned on the bare Ag surface about 1 nm laterally away from the molecule. The bottom row shows 2D TERS images obtained integrating the intensity of the greyed-out bands. Some pixilation is still visible corresponding to the finite step size (~0.16 nm) used to obtain these 2D intensity maps (23 × 23 pixels). On the bottom right an UHVSTM topography image of an individual H2TBPP molecule adsorbed flat at low temperature is shown. Reproduced with permission from Ref. [40].

Some reports demonstrated that the TERS resolution can even reach below the nanometer level [40]. The spectra and Raman images in Fig. 5.8 refer to a single porphyrin molecule deposited on an Ag substrate. A low-temperature STM under UHV conditions was used to minimize molecular diffusion on the flat surface and achieve atomic resolution in the STM topography images to be compared with the TERS images. The gap-mode STM configuration (Au tip/ Ag surface) is compulsory to have the highest field enhancement and to be sensitive to a single molecule. Still, it was further required to finely tune the LSP resonance of the tip-sample cavity with the energy of the Raman-scattered photons by controlling with Å precision the tip to sample distance. Note that the quadrilobe structure of individual porphyrin molecules is distinguished also in the topographic STM images; on the other hand, the TERS images on the bottom are indeed optical images recorded on a very narrow

131

132

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

wavelength window (corresponding to photons Raman scattered by a particular molecular vibration) within the visible range. The resolution is down to only a few Å in both kinds of images. Still, this submolecular optical resolution is not fully compatible with the common TERS image in which the lateral resolution is associated with the physical size of the apex and with the external electromagnetic field enhancement associated with the LSP excitation in nanostructure with dimensions 5 to 30 nm. A full rationalization of TERS at an atomic level is still missing, and more experimental evidence must be gathered. More advanced concepts have been invoked [41], some of which common to SERS: For example, smaller structures (with nanoscale roughness on the larger spherical apex) can be responsible for even higher field strength and field localization and/or for an additional “internal” enhancement (reminiscent of the chemical effect) present only for specific interactions between the tip and molecule at very short distances, possibly involving molecular tunneling and specific hybrid states of the molecule and clusters of metal atoms.

5.3 SERS/TERS Applications in PlasmonMediated Chemical Reactions

SERS and TERS are analytical techniques useful to chemically identify molecules surrounding the illuminated metallic nanoparticles supporting the LSP optical excitation. By adopting well-planned procedures, these techniques can be used to monitor in time the chemical reactions taking place at the metal surface, ideally as the spectral signature of the reactants transforms into the spectral signature of the products, or more generally by detecting spectral changes indicative of a surface process taking place. In these instances, the illuminated nanoparticles play the double role of Raman signal enhancer (sensing the environment) and promoter of chemical changes (modifying the environment). In both cases, the efficient excitation of LSP modes is the key step. To disentangle the two roles, two general strategies exist: In a number of reports, an illumination intensity threshold exists for triggering surface chemical process, likely related to the number of hot electrons generated and transferred to the adsorbate or to the required

SERS/TERS Applications in Plasmon-Mediated Chemical Reactions

temperature increase associated with the thermoplasmonic effect. The SERS characteristics can be first probed when illuminating below the threshold, to avoid or minimize the modification of the original molecular structure; then, laser pulses with intensity above the threshold value are applied, and finally the metal surface is again probed with reduced laser power. A humble precursor of this approach is the “burning” of a molecular dye layer under strong illumination. The molecules adsorbed at a metal substrate provide a recognizable pattern of Raman bands when illuminated at medium low intensity. If high laser powers are used, very strong and mostly random SERS/TERS fluctuations appear, signaling the tumultuous formation of many different and short-lived chemical species or intermediates (which can be difficult to identify due to band overlaps), until only amorphous carbon traces are left on the surface. This complex and very fast surface chemistry (involving also oxygen from air) is the result of the LSP excitation and the generation of highly energetic electrons near the surface. An example predating the interest in plasmon-enhanced chemical reaction is [42], which mainly focuses on the evaluation of the hottest SERS sites, their distribution and contribution to the overall SERS signal using an Ag film on nanospheres (FON) (Fig. 5.9). Molecules at the hottest sites are “damaged” when the local field gEincident ≥ Eth, where Eth is the threshold field, depending on the type of molecule, the wavelength, and the pulse duration and g is the field enhancement factor due to the LSP excitation. The molecules located where g is largest “burn” first. Spectrally, there is a loss of intensity (3050 cm−1 phenyl nCH) or broadening (near 1550 cm−1) of some of the original bands and the appearance of new features (near 2100 cm−1, carbonyl stretch resulting from the interaction of the photoproduct with oxygen). Anyhow, the used pulses (with energy in the range of 3 to 20.107 V/m) were not intense enough to damage the substrate and/or modify the magnitude of the SERS effect. More recently, studying the SERS/TERS spectral variations for increasing laser fluency, Sczerbinsky et al. [43] focused on the molecular cross-linking (formation of intermolecular C–C bonds) in a series of model aliphatic and aromatic thiol self-assembled monolayers (SAMs). The products are equaled to those observed upon X-ray or electron beam irradiation, supporting reaction mechanisms involving hot electrons with energies up to a few eVs

133

134

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

above the metal Fermi level generated upon visible light illumination. These are assimilated to the secondary electrons scattered by the metal substrate upon nominally much more energetic X-ray irradiation. Ruling out of thermal mechanisms is based on comparatively analyzing the products under thermal programmed desorption mass spectrometry (TPD-MS). These results support an alternative explanation for the origin of the well-documented (and often limiting) SERS/TERS spectral fluctuations, toning down the role of local temperature increase upon the illumination of the SERSactive substrate (“burning” of the molecular adlayer in the presence of ambient oxygen) and designate the surface release and further recombination of reactive molecular fragments, such as radicals formed in the presence of hot electrons, in an induced desorption process.

Figure 5.9 In this study, the original molecular layer (benzenethiol molecules BT, represented by straight lines) adsorbed on the Ag substrate is analyzed by SERS (upper row). Intense laser pulses of increasing power are then used to illuminate the sample, and photoproducts are formed and subsequently analyzed by SERS (bottom row). The spectra show the presence of new species, mostly localized in the gap between adjacent spheres (small black circles), where electromagnetic fields are stronger. Reproduced with permission from Ref. [42].

SERS/TERS Applications in Plasmon-Mediated Chemical Reactions

A second approach to study chemical reactions by means of SERS/TERS is based on adopting a dual-wavelength illumination: One wavelength is used to promote the chemistry, and a different wavelength is used to excite the SERS/TERS effect. In principle, this approach allows to follow in real time the chemical transformations. The two chosen wavelengths should fall within the same broad LSP resonance (up to ~100 nm wide) for the NP to actively participate in both mentioned roles. In this respect, note that the sites of maximal SERS effect might not be necessarily the same sites responsible for the strongest chemical reactivity. This is especially true for the thermoplasmonic effect, which mostly relies on the effects of the vigorous electron oscillation within the metal body (ultimately, joule heating) as opposed to the generation of strong electromagnetic fields outside the nanoparticles necessary for SERS [44]. Concerning reaction mechanisms involving hot electrons, not only their generation should be considered, which can be more or less related to the local field intensity, but also their lifetime and transfer to the adsorbate, which can also be site dependent. The field of plasmon-assisted chemical reaction is still dawning, and a number of experimental and theoretical issues need more investigations, accepting that concepts valid for a given chemical system (“reaction”) may not be transferable to a different system. The difficulty to generalize also lies on the specific way a molecule interacts with the LSP exciton. A molecular system that has been often investigated by both SERS and TERS is the dimerization of 4-aminothiophenol (4ATP, terminal amino group) and 4-nitrothiophenol (4NTP, terminal nitro group) SAMs to the same dimercaptoazobenzene (DMAB) product, presenting an N=N double bound, by oxidation and reduction, respectively [45]. All those molecules interact strongly with the metal surface via the S atoms, display large Raman cross sections (mostly due to the aromatic skeleton), and present rather clean spectral signatures. For the oxidation process involving two adjacent 4ATP molecules, the appearance of different, additional bands was initially not recognized due to newly formed chemical species, but rather to an exceptional surface enhancement of otherwise weak bands of b2 symmetry. Tian’s group and Sun’s group [46, 47] advanced the molecular dimerization process, depicted in Fig. 5.10, and the bands at 1140 cm−1, 1388 cm−1, and 1438 cm−1 were assigned to dCH and nNN

135

136

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

modes from the dimer at the Ag roughened surface. The scissoring of the double bond of the surface-attached DMAB molecules (thus a reduction back to two molecules of 4ATP) is also possible, and in this reverse case, a participation of hot electrons was invoked with H2 or H2O acting as the hydrogen source [48, 49]. Alternating proper oxidizing and reducing atmospheres, the molecular dimerization and scission may be cyclically repeated.

Figure 5.10 On the left, the dimerization of 4ATP (bottom row) to DMAB (upper row) and reverse process. The additional DMAB bands are evidenced in red. Reproduced with permission from Ref. [46]. On the right, the mechanism proposed by Xu et al. for the conversion of 4ATP to DMAB under irradiation and plasmonic heating, further involving oxygen as the electron acceptor (oxidant) and water molecules for the deprotonation of 4ATP (basic condition accelerates the process). Reproduced with permission from Ref. [48]. Note that the NN bond breaking follows a different mechanism involving hot electrons.

SERS/TERS Applications in Plasmon-Mediated Chemical Reactions

Figure 5.11 (Left) Schematics of the TERS setup used to monitor catalytic processes on metal surfaces at the nanoscale, in particular the reduction of 4NTP to DMAB. (Right) Time evolution of the TERS intensity at 633 nm excitation (waterfall plot) at 633 nm excitation before (top region, no change in the spectral feature of 4NTP) and after illumination at 532 nm (below white band). Two individual spectra from the waterfall plot: Spectrum (i) is taken at 90 s (original 4NTP SAM) and spectrum (ii) at 265 s. Reproduced with permission from Ref. [52].

The reduction process of two adjacent 4NTP molecules to DMAB on Au, Ag, and Cu SERS substrates illuminated at different wavelengths is documented in Refs. [50, 51]. The spectral evidence is the disappearance of the ns signal of the nitro group at 1331 cm−1 and the appearances of new bands at 1387 cm−1 and 1432 cm−1 having strong contributions from the nitrogen double bond in DMAB. Four hot electrons stemming from the LSP decay are

137

138

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

required for the reduction of two nitro groups and the formation of a DMAB molecule. This hot-electron-driven process has been also investigated by TERS [52, 53]. In Ref. [52], the authors use as semitransparent substrate (bottom illumination AFM-TERS) chemically synthesized nano-plates of Au dispersed on glass for the tethering of the 4NTP SAM and to profit from the gap mode also under bottom illumination with a large NA objective (Fig. 5.11). The nano-plates present an ultrasmooth surface comparable to the 111 terraces of bulk crystalline Au. Furthermore, in this work, the double wavelength approach is used, with 532 nm light used to trigger the photocatalytic activity and the less invasive (for this gap configuration) 633 nm light only for monitoring purposes.

Figure 5.12 Scheme of the LSP-assisted reduction of 4NTP to 4ATP on laserilluminated AgNP in the presence of AgCl. Holes h+ are left on the NP (3), which are captured by halide ions (4), leading to the precipitation of the Ag salt (5), which readily photodissociates under illumination (6). Reproduced with permission from Ref. [54].

An important and general observation, inferred from the reduction of 4NTP to 4ATP, was made by Xie and Schuckler [54]. When using Ag nanoparticle in the presence of AgCl in solution, the reaction proceeds also in the absence of the reducing agent BH4− ions. The halide ions play a decisive role compensating the holes left on the AgNPs after the hot electron transfer and reduction of 4NTP, via an oxidation half reaction and recycling of the plasmonically active Ag surface. As exemplified in Fig. 5.12, overall six electrons are

SERS/TERS Applications in Plasmon-Mediated Chemical Reactions

required to complete the reduction of one 4NTP molecule (one full cycle); the oxidation counterpart (steps from 4 to 7) can be globally written as h+ (hole left in the AgNP) + Cl-s Æ Cl·. The Cl· radical is actually formed (step 6) from the photodissociation of the poorly soluble AgCl, which precipitates on the NP. Generally speaking, whenever a hot electron mechanism is advanced, the fate of the simultaneously generated hot holes is also important and it can actually limit the whole chemical process, if these just recombine after a short time with the hot electrons: A suitable oxidation half reaction must coexist to sustain the reduction and balance the charge left of the NP. The work of Jain also underlines this point [55]. Tijunelyte et al. [56] employed Au nanocylinders of different diameters fabricated by electron beam lithography (EBL) as a tunable substrate for a thiol-ene “click” chemical reaction. 1,4-butanedithiol, bearing a double-bond terminal group, was initially self-assembled on the Au cylinders. The modified surface was then exposed to thiophenol in the liquid phase in the presence of a radical catalyzer (AAPH, active at temperature above 60°C or under UV illumination) and irradiated by 660 nm laser light (Fig. 5.13). The reaction kinetics were then monitored by SERS in real time with acquisitions every minute following the increase in the intensity of the aromatic ring signal (nC=C at 1572 cm−1), as the thiophenol molecules from the solution reacted with the allyl molecules grafted on the Au surface and were thus increasingly immobilized there. With the support of several negative control experiments (for example to exclude spontaneous adsorption of thiophenol on gold by thiol exchange) and by testing with different sizes of cylinders (110 nm, 140 nm, 200 nm in diameter), characterized by LSP resonances at different wavelengths, the authors prove that the illumination of the 110 nm AuNP (having the extinction maximum the closest to 660 nm) accelerates up to 20 times the reaction kinetics and that this effect is related to the LSP excitation. Indeed, the reaction time upon illumination at the used fixed wavelength is strongly dependent on the LSP position. Concerning the operating mechanism(s), both a thermoplasmonic effect, sufficient to raise the local temperature above 60°C, and a broadening or shifting of the electronic levels of the AAPH catalyzer (normally requiring UV excitation to initiate the radical cycle) upon interaction with the gold surface, allowing the absorption of less energetic photons in the visible region, are advanced.

139

140

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

Figure 5.13 Schematics of the click-chemistry reaction taking place at the surface of laser-illuminated Au nanocylinders. The radical initiator (2,2′-azobis(2methylpropionamidine) dihydrochloride (AAPH)) forms a thiyl radical, which then reacts with the terminal C=C bond of the surface molecules. On the right, normalized SERS intensities (1572 cm−1) versus reaction time for the 110 nm (black dot), 140 nm (red dot), and 200 nm (blue dot) nanocylinders. Reproduced with permission from Ref. [56].

SERS/TERS Applications in Plasmon-Mediated Chemical Reactions

In the following last example, Raman spectroscopy is not used, but the entire setup and the used probes are exactly those employed for gap-mode TERS using an STM (under UHV environment) [57]. The authors studied in real time by high-resolution STM, the photodissociation of the dimer molecule (CH3S)2 taking place at single-molecule level within the plasmonic cavity formed between a tunneling Ag tip and an Ag surface. The main message of this work is to demonstrate that this molecular dissociation proceeds by direct intramolecular excitation induced by the LSP (energy directly transferred from the LSP to the molecule, more precisely to a LUMO hybridized with the metal states but still presenting a large s*S-S character) and not by an indirect LSP mechanism involving LSP decay, generation of hot electrons, their transfer to form a transient negative ion (TNI), which finally decays to highly excited vibrational states (leading to dissociation) (see Fig. 5.14) as advanced for other plasmon-assisted photo-dissociations (O2 and H2). The evidence is based on two series of experiments: First, plotting the wavelength dependence of the plasmon-induced chemical reaction reveals more similarities to the “normal” photodissociation process (no tip, no LSP excitation), although taking place in a shifted energy range due to the involvement of LSP excitation. No energy threshold effect is observed, which would be expected for the transfer of hot electrons. They further compared real-time STM results for the plasmon-induced reaction (hypothesis of direct mechanism) and the same process taking place in the dark but under inelastic electron tunneling, in which the energy of the electrons can be controlled by the applied tunneling voltage and are shuttled through the nano-cavity, essentially “creating” ballistic hot electrons without illumination and excitation of the LSP modes (indirect pathway). The degree of hybridization between the molecular orbitals and the metal states (or more generally the electronic interaction of the molecule with the metal defined by their MO structure and DOS) plays a crucial role in the mechanism underlying the plasmoninduced chemical reactions.

141

142

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

Figure 5.14 (Above left) Schematic of the molecular photodissociation process induced by illuminating the nano-cavity formed between the tunneling Ag tip and the metal substrate. The red area symbolizes the localization region of the excited LSP gap mode. In the middle, high-resolution, UHV-STM images of an individual (CH3S)2 dimer, before and after dissociation (the tip was located at the position marked with a cross, the scale bar is 0.5 nm). (Bottom left) The tunneling current under constant height STM operation mode recorded during the photodissociation of a single molecule; at t = 4 s, laser light is switched on and at t = 18 s switched off. The current drop at t = 12 s is due to the change in the tunneling resistivity (i.e., tunneling distance) when the molecule is dissociated. The delay time tR depends on the local field intensity: stronger LSP excitation (for smaller gap distances), shorter tR. On the left, the two discussed mechanisms involving the hot electron transfer to form a TNI (indirect mechanism) and the direct pathway. Reproduced with permission from Ref. [57].

References 1. Fleischmann, M., P. J. Hendra, and A. J. McQuillan, Raman spectra of pyridine adsorbed at a silver electrode. Chemical Physics Letters, 1974. 26(2): 163–166. 2. Jeanmaire, D. L. and R. P. Van Duyne, Surface Raman spectroelectrochemistry: Part I. Heterocyclic, aromatic, and aliphatic

References

amines adsorbed on the anodized silver electrode. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, 1977. 84(1): 1–20.

3. Albrecht, M. G. and J. A. Creighton, Anomalously intense Raman spectra of pyridine at a silver electrode. Journal of the American Chemical Society, 1977. 99(15): 5215–5217. 4. Pemberton, J. E., et al., Surface enhanced Raman scattering in electrochemical systems: The complex roles of surface roughness. Applied Surface Science, 1988. 32(1): 33–56.

5. Wetzel, H., H. Gerischer, and B. Pettinger, Surface enhanced Raman scattering from silver-halide and silver-pyridine vibrations and the role of silver ad-atoms. Chemical Physics Letters, 1981. 78(2): 392– 397. 6. Kudelski, A. and J. Bukowska, Charge-transfer contribution to surfaceenhanced Raman scattering and surface-enhanced resonance Raman scattering of dyes at silver and gold electrodes. Chemical Physics Letters, 1996. 253(3): 246–250.

7. Wu, D.-Y., et al., Electrochemical surface-enhanced Raman spectroscopy of nanostructures. Chemical Society Reviews, 2008. 37(5): 1025–1041. 8. Yang, K.-H., Y.-C. Liu, and C.-C. Yu, Enhancements in intensity and stability of surface-enhanced Raman scattering on optimally electrochemically roughened silver substrates. Journal of Materials Chemistry, 2008. 18(40): 4849–4855. 9. Gao, P., et al., Surface-enhanced Raman scattering at gold electrodes: Dependence on electrochemical pretreatment conditions and comparisons with silver. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry, 1987. 233(1): 211–222.

10. Kudelski, A., et al., Characterization of the copper surface optimized for use as a substrate for surface-enhanced Raman scattering. Vibrational Spectroscopy, 1998. 16(1): 21–29. 11. Jacobse, L., et al., Correlation of surface site formation to nanoisland growth in the electrochemical roughening of Pt(111). Nature Materials, 2018. 17(3): 277–282.

12. Leung, L. W. H. and M. J. Weaver, Extending surface-enhanced Raman spectroscopy to transition-metal surfaces: Carbon monoxide adsorption and electrooxidation on platinum- and palladium-coated gold electrodes. Journal of the American Chemical Society, 1987. 109(17): 5113–5119.

143

144

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

13. Tian, Z.-Q. and B. Ren, Adsorption and reaction at electrochemical interfaces as probed by surface-enhanced Raman spectroscopy. Annual Review of Physical Chemistry, 2004. 55(1): 197–229.

14. Mrozek, M. F., Y. Xie, and M. J. Weaver, Surface-enhanced Raman scattering on uniform platinum-group overlayers: Preparation by redox replacement of underpotential-deposited metals on gold. Analytical Chemistry, 2001. 73(24): 5953–5960.

15. Maver, U., et al., Recent progressive use of atomic force microscopy in biomedical applications. TrAC Trends in Analytical Chemistry, 2016. 80: 96–111. 16. Seo, Y. and W. Jhe, Atomic force microscopy and spectroscopy. Reports on Progress in Physics, 2007. 71(1): 016101. 17. Alessandrini, A. and P. Facci, AFM: A versatile tool in biophysics. Measurement Science and Technology, 2005. 16(6): R65–R92.

18. Flater, E. E., et al., Towards easy and reliable AFM tip shape determination using blind tip reconstruction. Ultramicroscopy, 2014. 146: 130–143. 19. Pettinger, B., et al., Surface-enhanced and STM-tip-enhanced Raman spectroscopy at metal surfaces. Single Molecules, 2002. 3(5–6): 285– 294.

20. Chang, R. K. and T. E. Furtak, Surface Enhanced Raman Scattering, 1982. New York: Plenum Press. viii, 423 p. 21. Le Ru, E. C. and P. G. Etchegoin, Principles of Surface-Enhanced Raman Spectroscopy: And Related Plasmonic Effects, 1st ed, 2009. Amsterdam, Boston: Elsevier. xxiii, 663 p. 22. Langer, J., et al., Present and future of surface-enhanced Raman scattering. ACS Nano, 2020. 14(1): 28–117.

23. Pettinger, B., Light scattering by adsorbates at Ag particles: Quantum‐ mechanical approach for energy transfer induced interfacial optical processes involving surface plasmons, multipoles, and electron‐hole pairs. The Journal of Chemical Physics, 1986. 85(12): 7442–7451. 24. Furtak, T. E. and S. H. Macomber, Voltage-induced shifting of chargetransfer excitations and their role in surface-enhanced Raman scattering. Chemical Physics Letters, 1983. 95(4): 328–332.

25. Lombardi, J. R., et al., The effect of molecular structure on voltage induced shifts of charge transfer excitation in surface enhanced Raman scattering. Chemical Physics Letters, 1984. 104(2): 240–247.

26. Otto, A., The ‘chemical’ (electronic) contribution to surface-enhanced Raman scattering. Journal of Raman Spectroscopy, 2005. 36(6–7): 497–509.

References

27. Langer, J., S. M. Novikov, and L. M. Liz-Marzán, Sensing using plasmonic nanostructures and nanoparticles. Nanotechnology, 2015. 26(32): 322001. 28. Haynes, C. L., et al., Surface-enhanced Raman sensors: Early history and the development of sensors for quantitative biowarfare agent and glucose detection. Journal of Raman Spectroscopy, 2005. 36(6–7): 471–484. 29. Wessel, J., Surface-enhanced optical microscopy. Journal of the Optical Society of America B, 1985. 2(9): 1538–1541.

30. Webster, S., D. N. Batchelder, and D. A. Smith, Submicron resolution measurement of stress in silicon by near-field Raman spectroscopy. Applied Physics Letters, 1998. 72(12): 1478–1480. 31. Inouye, Y. and S. Kawata, Near-field scanning optical microscope with a metallic probe tip. Optics Letters, 1994. 19(3): 159–161.

32. Pettinger, B., et al., Tip-enhanced Raman scattering: Influence of the tip-surface geometry on optical resonance and enhancement. Surface Science, 2009. 603(10): 1335–1341.

33. Pettinger, B., et al., Tip-enhanced Raman spectroscopy: Near-fields acting on a few molecules. Annual Review of Physical Chemistry, 2012. 63(1): 379–399. 34. Schmid, T., et al., Nanoscale chemical imaging using tip-enhanced Raman spectroscopy: A critical review. Angewandte Chemie International Edition, 2013. 52(23): 5940–5954. 35. Mauser, N. and A. Hartschuh, Tip-enhanced near-field optical microscopy. Chemical Society Reviews, 2014. 43(4): 1248–1262.

36. Zhang, Z., et al., Tip-enhanced Raman spectroscopy. Analytical Chemistry, 2016. 88(19): 9328–9346.

37. Pettinger, B., et al., Nanoscale probing of adsorbed species by tipenhanced Raman spectroscopy. Physical Review Letters, 2004. 92(9): 096101. 38. Rendell, R. W. and D. J. Scalapino, Surface plasmons confined by microstructures on tunnel junctions. Physical Review B, 1981. 24(6): 3276–3294.

39. Anderson, N., et al., Nanoscale vibrational analysis of single-walled carbon nanotubes. Journal of the American Chemical Society, 2005. 127(8): 2533–2537. 40. Zhang, R., et al., Chemical mapping of a single molecule by plasmonenhanced Raman scattering. Nature, 2013. 498(7452): 82–86.

145

146

SERS and TERS and Their Applications in Organic Synthesis or Catalysis

41. Richard-Lacroix, M., et al., Mastering high resolution tip-enhanced Raman spectroscopy: Towards a shift of perception. Chemical Society Reviews, 2017. 46(13): 3922–3944.

42. Fang, Y., N.-H. Seong, and D. D. Dlott, Measurement of the distribution of site enhancements in surface-enhanced Raman scattering. Science, 2008. 321(5887): 388–392. 43. Szczerbiński, J., et al., Plasmon-driven photocatalysis leads to products known from E-beam and X-ray-induced surface chemistry. Nano Letters, 2018. 18(11): 6740–6749.

44. Lalisse, A., et al., Quantifying the efficiency of plasmonic materials for near-field enhancement and photothermal conversion. Journal of Physical Chemistry C, 2015. 119(45): 25518–25528. 45. Ren, X., et al., Recent advances in surface plasmon-driven catalytic reactions. RSC Advances, 2017. 7(50): 31189–31203.

46. Huang, Y.-F., et al., When the signal is not from the original molecule to be detected: Chemical transformation of para-aminothiophenol on Ag during the SERS measurement. Journal of the American Chemical Society, 2010. 132(27): 9244–9246. 47. Fang, Y., et al., Ascertaining p,p′-dimercaptoazobenzene produced from p-aminothiophenol by selective catalytic coupling reaction on silver nanoparticles. Langmuir, 2010. 26(11): 7737–7746. 48. Xu, P., et al., Mechanistic understanding of surface plasmon assisted catalysis on a single particle: Cyclic redox of 4-aminothiophenol. Scientific Reports, 2013. 3(1): 2997.

49. Sun, M., et al., Plasmonic scissors for molecular design. Chemistry: A European Journal, 2013. 19(44): 14958–14962.

50. Dong, B., et al., Substrate-, wavelength-, and time-dependent plasmonassisted surface catalysis reaction of 4-nitrobenzenethiol dimerizing to p,p′-dimercaptoazobenzene on Au, Ag, and Cu films. Langmuir, 2011. 27(17): 10677–10682.

51. Dong, B., et al., Is 4-nitrobenzenethiol converted to p,p′dimercaptoazobenzene or 4-aminothiophenol by surface photochemistry reaction? Journal of Raman Spectroscopy, 2011. 42(6): 1205–1206. 52. van Schrojenstein Lantman, E. M., et al., Catalytic processes monitored at the nanoscale with tip-enhanced Raman spectroscopy. Nature Nanotechnology, 2012. 7(9): 583–586.

53. Sun, M., et al., In-situ plasmon-driven chemical reactions revealed by high vacuum tip-enhanced Raman spectroscopy. Scientific Reports, 2012. 2(1): 647.

References

54. Xie, W. and S. Schlücker, Hot electron-induced reduction of small molecules on photorecycling metal surfaces. Nature Communications, 2015. 6(1): 7570. 55. Kim, Y., D. Dumett Torres, and P. K. Jain, Activation energies of plasmonic catalysts. Nano Letters, 2016. 16(5): 3399–3407.

56. Tijunelyte, I., et al., Nanoplasmonics tuned “click chemistry.” Nanoscale, 2016. 8(13): 7105–7112.

57. Kazuma, E., et al., Real-space and real-time observation of a plasmoninduced chemical reaction of a single molecule. Science, 2018. 360(6388): 521–526. 58. Prodan, E., et al., A hybridization model for the plasmon response of complex nanostructures. Science, 2003. 302(5644): 419–422. 59. McMahon, J. M., et al., Modeling the effect of small gaps in surfaceenhanced Raman spectroscopy. The Journal of Physical Chemistry C, 2012. 116(2): 1627–1637. 60. Hao, F., et al., Plasmon resonances of a gold nanostar. Nano Letters, 2007. 7(3): 729–732.

61. Nordlander, P., et al., Plasmon hybridization in nanoparticle dimers. Nano Letters, 2004. 4(5): 899–903.

147

Index

Abbe’s principle 125 absorption 11–16, 21, 24, 30–31, 58, 114–115, 139 absorption spectroscopy 30 acetaldehyde 55, 97 acetone 55–56, 65 acetophenone 50, 68 acid 75 4-azidobenzoic 80 4-mercaptophenylboronic 75 aryl boronic 71 corresponding carboxylic 57 formic 76 phenylboronic 74 adhesiveness 44, 46 adsorbate 118, 132, 135 adsorption 52, 59, 78, 113, 122, 128, 139 AFM see atomic force microscope AFM platform 42, 44 AFM probe 40, 124 Ag nanoparticle (AgNP) 19–21, 55–57, 61, 63, 102, 138–139 AgNP see Ag nanoparticle alcohol 50–51, 56, 61, 65, 81–83 aliphatic 55 aromatic 50, 52, 55–56 benzyl 50, 52, 55–56, 75, 78, 82–83 benzylic 51 cinnamyl 65 furfuryl 68 phenethyl 50, 52 secphenethyl 51 aldehyde 50, 55–56, 65, 77, 82, 104 alkane 57–58 alkene 57, 61, 77, 80, 97 alkyne 77–78

amine 50, 53, 77, 81–82 amplification 116, 118 aniline 60, 62, 64, 78, 82 approximation 15, 116, 125 aromatic compound 63, 74 aryl halide 71, 75, 101 atomic force microscope (AFM) 36–37, 39–46, 124–126 AuNP see gold nanoparticle azobenzene 59–61 azoxybenzene 60

benzene 56–57 benzylamine 53, 79 bipyridine 99, 102–103 bottom illumination 126–127, 129, 138 cantilever 37–42 catalysis enantioselective 98 heterogenous 98 hydrolytic 99–100 photo-assisted 98 plasmon-enhanced 64 supported molecular 98 catalyst 50, 56, 58–59, 63, 67–68, 71–72, 79, 98–99, 101, 104 classical 71 copper-bipyridine 99 heterogeneous 68, 70 nanorattle 67 photoredox 79 selective 99 catalytic activity 56, 58–59, 61–63, 71, 77–78, 117 CB see conduction band charge transfer (CT) 63, 118–119, 121

150

Index

chemical enhancement 118–121 chemical sensitivity 46, 125 collision 2–3, 31 colloid 33, 106, 118 conduction band (CB) 2, 79, 118 copper 68, 80, 83, 99, 101–102, 104, 106, 123 CT see charge transfer curcumin 63

dark conditions 55, 59, 63, 70–71, 73–75, 106 delocalized surface plasmon (DSP) 1, 5–6, 9, 12 dielectric constant 3–6, 8, 11, 16, 19, 21–22, 122 dielectric medium 5–6, 8–11, 13 dimercaptoazobenzene (DMAB) 61, 135–137 dimerization 61–62, 135–136 DMAB see dimercaptoazobenzene Drude–Lorentz model 2–3 Drude model 2–3, 5, 7 DSP see delocalized surface plasmon dye 65, 69, 113, 118, 128

EBL see electron beam lithography EELS see electron energy loss spectroscopy electron beam lithography (EBL) 33, 46, 139 electron energy loss spectroscopy (EELS) 34–35 electron probe resonance (EPR) 60 emission 31–32, 34, 66 energy 1, 5, 31–32, 35–36, 68–69, 98, 114–116, 118–119, 131, 133, 141 electron 5 electronic 116 optimal excitation 119 orbital 120 photon 24, 114

plasma oscillation 5 plasmon absorption bands dissipating 101 thermal 55 visible-light 80 EPR see electron probe resonance ethanol 19–20, 50, 55, 74, 106 evanescent wave 9–10 excitation 1, 3–6, 22, 24, 50–51, 66–67, 77, 99, 101, 113–114, 118–119, 124, 127, 137 intramolecular 141 optical 113, 132 plasmon-induced 80 resonant 114 excitation frequency 3–4 excitation light 6, 9–11 excitation polarization 21 excitation wavelength 11, 14–15, 17, 19, 21, 61 extinction 13–14, 30, 46, 72 extinction spectrum 12–13, 19, 22–23, 29–31 Fermi–Dirac distribution 2 fluorescence 24, 66–67, 112–113, 116, 128 force 3, 36, 38–39 atomic 2, 36–39, 41, 44 elastic 3 viscosity 3 force constant 38, 40–41 gap 12, 21, 117, 126–127, 134 gold nanoparticle (AuNP) 50–52, 55–57, 59, 61–63, 65–67, 69–70, 76–81, 83, 99, 101, 104, 106 grafting 45, 102, 106 grating 9, 34

hot carriers 53, 58 hot electrons 24, 55, 58–59, 63, 74, 76, 79–81, 132–136, 139, 141

Index

hot holes 55, 76, 81, 139 hot spots 21, 114, 116–117, 120, 122–123, 126, 129 hydroamination 78 hydroazobenzene 60 hydrogenation 61, 65, 68, 97 hydrolysis 99, 102 hydroxylamine 66–67

illumination 30, 72, 114–115, 126–127, 133–135, 137–139, 141 interface 1, 5–6, 8–11, 50, 64, 113, 119 ion 2–3, 138 ionization 31–32, 34 irradiation 51–53, 55, 57, 59, 63–65, 67–68, 71, 73–76, 80–81, 104, 106–107 laser 46, 78, 104 light 56, 58–59, 63, 67, 71, 73–75, 78, 80 solar-light 57 ultraviolet 68 visible-light 55, 59, 64–65, 72, 74, 82 isopropanol 55–56, 61 ketone 61, 65, 104

laser argon 104 green 99 pulsed 51, 66–67, 70 laser diode (LD) 37, 40–41 laser interference lithography (LIL) 33 laser light 45–46, 68, 123, 125, 139, 142 lateral resolution 41, 43–44, 126, 130, 132 LD see laser diode lifetime imaging microscopy 52 ligand 99, 101–102, 104–106

LIL see laser interference lithography localized surface plasmon (LSP) 1–2, 5–6, 9, 11–21, 30, 98, 113, 115–116, 120–121, 132, 141 localized surface plasmon resonance (LSPR) 12, 21–22, 49, 55, 80–81, 98–99 LSP see localized surface plasmon LSP excitation 115–116, 118, 120, 124–125, 132–133, 139, 141–142 LSP mode 46, 114–115, 132, 141 LSPR see localized surface plasmon resonance LSP resonance 12, 45, 114–116, 120, 123–124, 131, 135, 139 LSP resonance shift technology 121 Maxwell equation 6, 13, 113 mechanism 51–55, 57, 60, 63, 67, 76, 79, 82–83, 99–100, 120–121, 136, 139, 141–142 electromagnetic 120 hot electron 139 metal-mediated hydrolytic 99 thermal 134 MEF see metal-enhanced fluorescence metal 1–2, 5–6, 8–13, 16–17, 21–22, 59, 75, 99, 113, 115–116, 118–121 bulk 5 buried 122 coinage 122, 125 free electron 121 inactive 122 ligand-stabilized transition 98 non-noble 68 non-plasmonic 68 substrate 113 transition 97

151

152

Index

metal-enhanced fluorescence (MEF) 24, 116 metal nanoparticle 49–50, 98, 113, 115 metal substrate 114, 128, 133–134, 142 methanol 50, 56, 69, 82–83 Mie coefficient 15 Mie model 15 Mie theory 13, 23 nanocatalyst 59, 71, 73–74, 81, 101–102, 106–107 nano-cavity 126, 129, 141–142 nanocrystal 58–59 nanocylinder 12, 81, 139–140 gold 12, 80 nanogap 75, 117, 122–123, 126 nanoheater 24, 67 nanoparticle (NP) 11, 13, 15–19, 21–22, 24, 30–31, 49–50, 52, 55–57, 102, 115, 117, 122–123, 135, 138–139 nanorod 12, 45–46, 53, 55, 64, 72, 117, 123 gold 33, 35–36, 46, 79 nanosphere 12–14, 16–19, 22–23, 133 nanostructure 2, 11–13, 33–34, 36, 44, 46, 58, 113, 117–118, 120, 132 hybrid 61, 80 metallic 12, 33 nearfield Raman imaging 130 nitroaromatics 59, 61, 64 nitrobenzene 60, 64 nitro group 61–62, 107, 135, 137–138 nitrophenol 62, 64 NP see nanoparticle

optical process 111–112, 114–116, 124 oscillation 1–4, 12, 39, 116

coherent 114 electron 22, 117, 127, 135 plasma 1 oxidation 50, 52, 55–58, 78, 97, 122, 135, 138–139 oxide 41, 50, 55, 57, 59, 61, 71–72, 77, 79, 81 oxygen 50, 52, 56, 79, 133–134

palladium 53, 71–74, 97 PCA see principal component analysis phenylacetylene 67, 77–78 phosphate ester 99–100 photocatalytic activity 55, 68, 77, 138 photocatalytic oxidation 55, 57 photodissociation 139, 141–142 photon 1, 13, 32, 34, 58, 114–116, 119, 139 photothermal effect 69 piezo elements 37, 41, 125 plasmon 1–2, 5, 7–11, 13, 22, 24, 31, 35, 115 plasmon excitation 7, 9–10, 15, 22, 24, 46, 75–76, 123 plasmonic catalyst 53, 55–56 plasmonic heating 76, 136 plasmonic properties 29, 34, 74 plasmonic substrate 46, 123 plasmon resonance 12–13, 16, 18–19, 21–22, 35, 55, 98, 114, 122 p-nitroaniline 60–61 polarizability 4, 16–19 ellipsoidal 18 molecular 118 principal component analysis (PCA) 36, 107 probe 36, 39–40, 42, 75, 123, 141 process 13, 15, 22, 24, 30, 50, 114–116, 119, 136, 141 bond-forming 101 charge-transfer 64

Index



chemical 132, 139 emission 32 extinction 15–16 hot-electron-driven 138 induced desorption 134 molecular dimerization 135 molecular photodissociation 142 normal photodissociation 141 oxidation/amidation 78 photonic 24 photon-scattering 119 physical 22 Raman 116, 118 redox 98 single-photon absorption 116 protocol 102, 104, 126

radiation 3, 17, 30, 74, 113 Raman bands 127, 133 Raman cross section 113, 118 Raman efficiency 112 Raman imaging 123, 130–131 Raman scattering 112, 114–115, 119, 124, 129–130 Raman signal 107, 112, 114, 116–117, 119–121, 123–124, 127–129, 132 Raman spectroscopy 46, 75, 111–112, 122, 124, 128, 141 Rayleigh approximation 17–19 Rayleigh scattering 115 reaction 50, 52–53, 55–57, 61, 64, 67–68, 70–81, 98–99, 101, 104, 106–107, 135, 138–139 amidation 78 bond-forming 104 catalytic 52, 61, 98–99, 104 chemical 46, 61, 77, 81, 132–133, 135, 139, 141 click-chemistry 80, 140 Friedel–Crafts 81 Henry’s 104–107 industrial-scale 97



intermolecular 77 oxidation/amidation 79 photocatalytic 49, 52, 63 plasmon-enhanced 75 plasmon-induced 141 protodeboronation 75–76 radical 68 Suzuki–Miyaura cross-coupling 53 Suzuki–Miyaura 71–75 thiol-ene click 80 Ullmann 101, 104 reaction mechanism 60, 133, 135 reactivity 67, 74–75, 80 redshift 15, 17–19, 22 reducing agent 61, 63–64, 67, 77, 80 reduction 17, 31, 57, 59–67, 74, 77, 101, 135–139 resazurin 65–67 resonance 4, 12–13, 15–19, 21, 31, 46, 60, 113, 115, 118–119, 127 resonance position 16–17, 19 resorufin 66–67 SAM see self-assembled monolayer scanning electron microscopy 31 scanning probe microscopy (SPM) 36, 124 scanning transmission electron microscopy (STEM) 32, 35 scanning tunneling microscope (STM) 36, 124–126, 131, 141 scattering 12–13, 15–16, 30–31, 114 inelastic 116, 120 phonon 80 SEIRA see surface-enhanced IR absorption selective hydrogenation 59, 62, 64–65 selective oxidation 50, 57 self-assembled monolayer (SAM) 75, 133, 135, 137–138

153

154

Index

SERS see surface-enhanced Raman scattering SERS effect 113, 122–123, 133, 135 SERS signal 117–118, 122–123, 133 SERS substrate 33, 121–123, 129 signal 32, 34, 37–38, 107, 112, 122, 130, 137, 139 error 38, 46 non-electronic 34 photodiode 37 spectroscopical 24 unenhanced far-field 129 Sommerfeld model 2 SPM see scanning probe microscopy SPP see surface plasmon polariton SPR see surface plasmon resonance STEM see scanning transmission electron microscopy STM see scanning tunneling microscope substrate 22, 33, 74–76, 79, 98, 123, 125–126, 129, 133 conductive 126 semitransparent 138 transparent 126 tunable 139 surface 5, 9, 11–13, 31, 34–42, 44, 55, 57–59, 68–69, 80–81, 112–114, 116–118, 120–128, 132–133, 139–140 archetypical 122 catalytic 117 conductive 127 dielectric 127 flat 42, 129, 131 illuminated 115 nanostructured 115, 122 planar 113 roughened 122, 136

soft 39 ultrasmooth 138 surface-enhanced IR absorption (SEIRA) 24, 116 surface-enhanced Raman scattering (SERS) 24, 33, 57, 60–61, 102, 106, 112, 114, 116–118, 121–123, 125, 127, 129, 132, 134–135 surface plasmon 1, 5–10, 12, 36, 58 surface plasmon polariton (SPP) 1, 6–10 surface plasmon resonance (SPR) 11, 36, 49–82, 98–99 Suzuki–Miyaura coupling 72, 74 Suzuki–Miyaura cross-coupling 71, 73–74

technique 29, 31, 36, 44, 51–52, 66, 69, 121, 124, 132 analytical 111, 132 characterization 42 etching 39 fluorescence-based 46 lithographic 33 nano-structuring 122 nearfield probe 118 optical deflection 41 quantitative analytical 122 SNOM 124 surface analysis 34 TERS see tip-enhanced Raman spectroscopy thermoelectric effect 32 thermoplasmonic effect 133, 135, 139 TiO2 50, 55–57, 59, 63, 74, 78, 102 tip 31, 36–44, 46, 118, 124–132, 141–142 tip apex 37, 42, 124–127 tip-enhanced Raman spectroscopy (TERS) 75, 112–142

Index

TIRFM see total internal reflection fluorescence microscopy TNI see transient negative ion total internal reflection fluorescence microscopy (TIRFM) 52

transient negative ion (TNI) 141–142

vacuum 4, 7, 14–16, 19–20, 23, 31 Young’s modulus 44

155